+ All Categories
Home > Documents > Modeling and simulation of photocatalytic degradation of ...

Modeling and simulation of photocatalytic degradation of ...

Date post: 02-Jan-2022
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
162
School of Engineering Sciences Faculty of Mechanical Science and Engineering Institute of Materials Science, Chair of Material Science and Nanotechnology Modeling and simulation of photocatalytic degradation of organic components in wastewater Hagen Eckert Born on: 28.03.1988 in Greifswald Dissertation to achieve the academic degree Doktoringenieur (Dr.-Ing.) Referees Prof. Dr. Gianaurelio Cuniberti Prof. Dr. Cormac H. Toher Prof. Dr. Hans-Peter Wiesmann Submitted on: 06.08.2020 Defended on: 15.12.2020
Transcript
Page 1: Modeling and simulation of photocatalytic degradation of ...

School of Engineering Sciences Faculty of Mechanical Science and Engineering

Institute of Materials Science, Chair of Material Science and Nanotechnology

Modeling and simulation ofphotocatalytic degradation oforganic components inwastewaterHagen EckertBorn on: 28.03.1988 in Greifswald

Dissertationto achieve the academic degree

Doktoringenieur (Dr.-Ing.)

Referees

Prof. Dr. Gianaurelio CunibertiProf. Dr. Cormac H. ToherProf. Dr. Hans-Peter Wiesmann

Submitted on: 06.08.2020

Defended on: 15.12.2020

Page 2: Modeling and simulation of photocatalytic degradation of ...
Page 3: Modeling and simulation of photocatalytic degradation of ...

Statement of authorship

I hereby certify that I have authored this Dissertation entitled Modeling and simula-tion of photocatalytic degradation of organic components in wastewater independentlyand without undue assistance from third parties. No other than the resources and

references indicated in this thesis have been used. I have marked both literal and ac-

cordingly adopted quotations as such. There were no additional persons involved in

the intellectual preparation of the present thesis. I am aware that violations of this

declaration may lead to subsequent withdrawal of the degree.

Dresden, 06.08.2020

Hagen Eckert

Page 4: Modeling and simulation of photocatalytic degradation of ...

EST AD ASTRA MOLLIS E TERRIS VIA - Seneca

Page 5: Modeling and simulation of photocatalytic degradation of ...

Acknowledgment

During my time at the TU Dresden, I had the pleasure of meeting a wonderful group

of people. The thesis as it stands now would not have been possible without them.

Foremost, I would like to thank Prof. Cuniberti for providing me the opportunity to

explore the numerous facets of computational material science. His guidance and

support throughout the years helped me to become a better scientist. Through his re-

lentless work, I was able to learn in a well-equipped research environment with brilliant

colleagues from all over the world.

I would like to thank Prof. Toher and Prof. Wiesmann for their interest in my research

and willingness to review this thesis.

Many people provided me with advice and took time out of their days to discuss my

questions and ideas. While not all of them can be mentioned by name, I would like to

point out Alexander, Arezoo, Rafael, Annegret, and Giovanni. I have to especially em-

phasize Manfred for helping me to keep my focus, and for his sound scientific inputs. I

would like to express my gratitude towards Manuela, Grit, and Sylvi, for being the back-

bone of our research. Without you, we would have gotten lost in the administrative

jungle long ago. While I took some detours, Sara, Nga, and Thomas provided me the

motivation to follow suit. I am thankful not just for the pleasant time we spent at work

but also for their friendships.

A crucial aspect of my time as PhD student was the opportunity to visit outstanding

research groups around the world. I would like to express my thanks to Prof. Clementi,

Prof. Persson, Prof. Pasquali, Prof. Lanceros-Mendez, and Prof. Shirai for being excel-

lent hosts. Here I also thank Scott and James, my roommates of choice, to guiding me

through the urban wilderness of Houston.

While research can be challenging at times, my friends and family kept me sane. Thank

you all for the big and small adventures we concluded and for the ones yet to come. I

have to formally apologize to Anna, Julia, Oliver, Philipp, Robert, and Sophia for making

them proofread parts of this thesis or to be beta-testers. Your sacrifices are much

appreciated. Numerous people accompany me on my path to this point in life. While

I sadly have to admit that I lost touch with some, all of them had an impact on me.

Thank you all for being part of this journey!

Page 6: Modeling and simulation of photocatalytic degradation of ...

Contents

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

Kurzfassung . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

Constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

1. Motivation 15

2. Introduction 192.1. Modeling and simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2.2. Heterogeneous photocatalysis . . . . . . . . . . . . . . . . . . . . . . . . . 22

2.2.1. History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2.2.2. Semiconductor band structure . . . . . . . . . . . . . . . . . . . . 23

2.2.3. Interface between a semiconductor and a redox electrolyte . . . 26

2.3. Photocatalytic material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

2.3.1. Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

2.3.2. Titanium dioxide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

2.3.3. Zinc oxide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

2.4. Light sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

2.4.1. Solar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

2.4.2. Fluorescent tubes and mercury-vapor lamps . . . . . . . . . . . . 42

2.4.3. Light-emitting diodes . . . . . . . . . . . . . . . . . . . . . . . . . . 43

2.4.4. Organic light-emitting diodes . . . . . . . . . . . . . . . . . . . . . 44

3. Materials and methods 453.1. Analytic methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.1.1. Nanoparticle characterization . . . . . . . . . . . . . . . . . . . . . 45

3.1.2. Ultraviolet-visible absorption spectrometry . . . . . . . . . . . . . 47

3.1.3. SPE-HPLC-MS/MS . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

3.1.4. Non-purgeable organic carbon . . . . . . . . . . . . . . . . . . . . 49

3.2. Experimental investigations . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

3.2.1. Model substances . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

3.2.2. Adsorption-desorption . . . . . . . . . . . . . . . . . . . . . . . . . 51

3.2.3. Photocatalytic degradation . . . . . . . . . . . . . . . . . . . . . . . 52

3.2.4. Wastewater treatment plant effluent . . . . . . . . . . . . . . . . . 53

6

Page 7: Modeling and simulation of photocatalytic degradation of ...

3.3. Modeling approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

3.3.1. Single organic species model . . . . . . . . . . . . . . . . . . . . . 58

3.3.2. Multiple organic species model . . . . . . . . . . . . . . . . . . . . 63

3.4. Model implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

3.4.1. Development objectives . . . . . . . . . . . . . . . . . . . . . . . . 72

3.4.2. Molecule parameters . . . . . . . . . . . . . . . . . . . . . . . . . . 73

3.4.3. Solving the differential equation system . . . . . . . . . . . . . . . 73

3.4.4. Fit to experimental results . . . . . . . . . . . . . . . . . . . . . . . 75

3.4.5. Availability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

4. Results and discussion 774.1. Nanoparticle properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

4.2. Adsorption-desorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

4.3. Photocatalytic degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

4.3.1. Single organic species model . . . . . . . . . . . . . . . . . . . . . 81

4.3.2. Multi organic species model . . . . . . . . . . . . . . . . . . . . . . 83

4.4. Wastewater treatment plant effluent . . . . . . . . . . . . . . . . . . . . . 89

4.4.1. Influence of effluent . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

4.4.2. Degradation of pharmaceuticals in the effluent . . . . . . . . . . . 91

5. Conclusions 95

Appendix 99A. Analytical solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

B. Effluent pharmaceuticals concentrations . . . . . . . . . . . . . . . . . . . 103

C. pdom handbook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

List of figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

List of tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

7

Page 8: Modeling and simulation of photocatalytic degradation of ...

Abstract

Organic pollutants are discharged into the water cycle at many stages in our daily

lives. Conventional wastewater treatments are ineffective in the removal of some

of them, especially clearing pharmaceuticals. Photocatalytic degradation utilizing cat-

alytic nanosuspensions under ultraviolet irradiation represents an efficient method

to reduce those organic components in the wastewater. While the general concept

of photocatalytic water purification is well established, a descriptive and easy to use

model of the essential processes was missing. Such a model is critical to ensure the

systematic comparability of experimental results and supports process optimization.

This work presents a modeling approach to simulate the involved kinetic processes

based on the Langmuir–Hinshelwood mechanism. Further, the fundamental model

is extended to include the formation of intermediate organic components. This ex-

tension uses either an incremental degradation mechanism or a fragmentation based

mechanism, that can include excess bonds. The simulated concentration evolution of

intermediates, as well as the evolution of the total organic carbon, are discussed for dif-

ferent model assumptions concerning their desorption rates from the photocatalyst

surface. The model parameters were estimated from comparison with experimental

findings. Basic experiments were performed using the antibiotic ciprofloxacin, and

the dye methylene blue as organic compounds and titanium dioxid and zinc oxide as

photocatalytic materials. Furthermore, the application of the model to more complex

systems is shown by the photocatalytic degradation of 14 pharmaceuticals in wastew-

ater treatment plant effluent. Following successful evaluation of this model, it was

implemented in an open-source software package to enable a wider adoption and a

sound foundation for further developments.

8

Page 9: Modeling and simulation of photocatalytic degradation of ...

Kurzfassung

Organische Schadstoffe werden in vielen Phasen unseres täglichen Lebens in den

Wasserkreislauf eingeleitet. Die herkömmliche Abwasserbehandlung ist nicht zur ef-

fektiven Entfernung einiger dieser Stoffe, insbesondere von Arzneimitteln, geeignet.

Die Fotokatalyse basierend auf der Suspension von katalytischen Nanopartikeln und

ultraviolettem Licht stellt eine effiziente Methode dar, um diese organischen Stoffe

im Abwasser zu reduzieren. Während das allgemeine Konzept der fotokatalytischen

Wasserreinigung gut etabliert ist, fehlte ein beschreibendes und einfach anwendbares

Modell der wesentlichen Abbauprozesse. Ein solches Modell ist entscheidend, um ex-

perimentelle Ergebnisse systematisch vergleichen zu können, und stellt eine wertvolle

Hilfe bei der Optimierung von Prozessen dar. Diese Arbeit präsentiert einen Model-

lierungsansatz zur Simulation der kinetischen Prozesse basierend auf dem Langmuir-

Hinshelwood-Mechanismus. Dieses Grundmodell wurde erweitert, um auch die Bil-

dung von organischen Zwischenprodukten zu beschreiben. Diese Erweiterungen ba-

sieren entweder auf einem inkrementellen oder einen fragmentierenden Abbaume-

chanismus, der durch das Einbinden von überschüssigen Bindungen ergänzt werden

kann. Die simulierte Konzentrationsentwicklung von Zwischenprodukten sowie die Ent-

wicklung des verbleibenden organischen Kohlenstoffes werden für verschiedene Mo-

dellannahmen bezüglich ihrer Desorptionsraten von der Photokatalysatoroberfläche

diskutiert. Die Modellparameter wurden aus dem Vergleich mit experimentellen Er-

gebnissen ermittelt. Grundlegende Experimente wurden unter Verwendung des Anti-

biotikums Ciprofloxacin und des Farbstoffs Methylenblau als Beispiele für organische

Verbindungen und Titandioxid und Zinkoxid als fotokatalytische Materialien durchge-

führt. Darüber hinaus wird die Anwendbarkeit des Modells auf komplexere Systeme

durch den Vergleich mit dem fotokatalytischen Abbau von 14 Medikamenten im Ab-

fluss von Kläranlagen demonstriert. Nach der Evaluierung des Modells wurde es in ein

Open-Source-Softwarepaket implementiert, um eine breitere Anwendung zu ermögli-

chen und eine solide Grundlage für weitergehende Entwicklungen zu schaffen.

9

Page 10: Modeling and simulation of photocatalytic degradation of ...

Symbols

Some symbols that are not central to this thesis are not listed in the following table,

but instead directly explained in the text near to their occurrence. Numbers in curly

brackets refer to chemical reactions.

Symbol Description Definition

α absorption coefficient 2.8

β excess bond split 3.40

δp penetration depth 2.13

ε molar extinction coefficient 3.6

εe energy state

η viscosity

κ adsorption prefactor (m) 3.29

μp hole mobility

ν photon frequency

ν0 pre-exponential factor (s−1) 3.29

φ quantum yield

Φ effective quantum yield 2.17

τp hole life-time

Θ surface coverage 3.14

A absorbance 3.6

as specific surface area (m−1) 3.9

Am surface area covered per molecule (m2) 3.14

A(ad) adsorbed organic molecule {8}

A(aq) organic molecule in solution {8}

b excess bond count {11}

CA(ad),0 initial concentration of adsorbed organic molecules (m−2)

CA(ad) concentration of adsorbed organic molecules (m2)

CA(aq),0 initial concentration of organic molecules in solution (m−3)

CA(aq) concentration of organic molecules in solution (m)

CA(ad),∞ equilibrium concentration of adsorbed organic molecules (m−2) 3.20

CA(aq),∞ equilibrium concentration of organic molecules in solution (m−3) 3.20

CBET BET parameter 3.1

10

Page 11: Modeling and simulation of photocatalytic degradation of ...

Symbol Description Definition

COx concentration oxidizing species

CRed concentration reduction species

D diffusion coefficient

E excess bond {11}

EF Fermi level

Ebend potential difference band bend

Edes desorption energy (J) 3.29

EF,redox redox fermi level 2.5

Eg bandgap energy

FF Fermi function 2.1

FK,M Kubelka-Munk remission function 3.5

I transmitted intensity 3.6

I0 photon flux (m−2 s−1)

Ii incident intensity 3.6

jads adsorption flux (m−2 s−1)

jdes desorption flux (m−2 s−1)

jph photocurrent density

jreac reaction flux (m−2 s−1)

K absorption coefficient 3.2

kads adsorption rate constant (m s−1) 3.11

kapp apparent rate constant (s−1) 3.22

kdes desorption rate constant (s−1) 3.11

kreac reaction rate constant (s−1) 3.11

Lp hole diffusion length 2.15

M(aq) mineralized components {8}

n molecule size index

Ns density of donor/acceptor states

NA molecule count organic species

na specific amount adsorbed (mol g−1) 3.1

nm specific monolayer amount of adsorbate (mol g−1) 3.1

p pressure of the adsorptive in equilibrium with the adsorbate (Pa) 3.1

p0 saturation vapour pressure of the adsorptive (Pa) 3.1

11

Page 12: Modeling and simulation of photocatalytic degradation of ...

Symbol Description Definition

R reflectance 3.4

S scattering coefficient 3.2

T temperature

TL light transmittance 3.6

U potential 2.3

U◦ standard potential

U◦redox

standard reduction potential

V volume

Wdl depletion layer width 2.14

z number of electrons

Constants

This work uses CODATA1 recommended values for fundamental physical constants [1].

The table below lists the values for the ones used.

Symbol Description Value

ε0 electric constant 8.854188 • 10−12 Fm−1

e elementary charge 1.602177 • 10−19 C

F Faraday constant 9.648533 • 104 Cmol−1

h Planck constant 6.626070 • 10−34 J s

kB Boltzmann constant 8.617330 • 10−5 eVK−1

NAvo Avogadro constant 6.022141 • 1023 mol−1

R molar gas constant 8.314460 J(mol K)−1

1Committee on Data for Science and Technology of the International Science Council

12

Page 13: Modeling and simulation of photocatalytic degradation of ...

Abbreviations

Abbreviation Description

AOP advanced oxidation processes

BET Brunauer-Emmett-Teller method

CB conduction band

HPLC high-performance liquid chromatography

IUPAC International Union of Pure and Applied Chemistry

LED Light-emitting diodes

LH Langmuir-Hinshelwood

LLoQ lower limit of quantification

LSODE Livermore Solver for Ordinary Differential Equations

MS/MS tandem mass spectrometry

NPOC non-purgable organic carbon

OLED organic light-emitting diodes

RHS right-hand side

SEM scanning electron microscope

SPE solid-phase extraction

TOC total organic carbon

UV ultraviolet

VB valence band

WWTP wastewater treatment plant

13

Page 14: Modeling and simulation of photocatalytic degradation of ...
Page 15: Modeling and simulation of photocatalytic degradation of ...

1. Motivation

Many pollutants, such as heavy metals, dyes, and pharmaceuticals, are discharged

into the water cycle at many stages in our daily lives. Pharmaceuticals are mostly in-

troduced in the sewage system through excretion of unmetabolized compounds af-

ter medical use or inappropriate disposal [2, 3, 4, 5], and reach the local wastewater

treatment plants (WWTPs). However, conventional WWTPs are not designed to treat

water that is polluted with trace levels of various pharmaceuticals. The conventionally

applied treatments are ineffective in their removal, in particular, due to interactions

of the biological cleaning stages with the pharmaceuticals [6, 7]. Consequently, they

reach the aquatic system and can be found in the surface and groundwater [8, 9], soil

and sediments [9, 10], and even in drinking [11, 12] and tap water [9, 13]. Commonly,

extreme diluted medicinal drugs, in the range of ng to µg per liter, do not present

acute toxic effects on aquatic organisms. Nevertheless, concerns have been raised

for chronic exposure, due to their continuous input into the environment, acting as

slightly persistent pollutants [3, 5, 14]. This problem is amplified through the high level

of pharmaceuticals that are used nowadays. The median non-hospital pharmaceuti-

cal consumption per capita in 2008 for high-income countries was 1042 SU2 [15]. This

subscription volume was a growth of over 18% compared to the year 2000. In par-

ticular, antibiotics represent a severe problem because of their toxicity to microflora

and -fauna, and of the possible development of antibiotic-resistant microorganisms

[16, 17, 18]. Global over- and misuse of antibiotics has intensified that problem and

resulted in steadily increasing concentrations in wastewater [19]. With nearly eight-

times higher consumption compared to low-income countries, rich countries need to

intensify their efforts to lower the over- and misuse of pharmaceuticals products and

2“One IMS Standard Unit equals one tablet, one capsule, one suppository or pessary, one pre-filled

syringe/cartridge, pen, vial or ampule, one dose of an inhaled medicine or 5mL of an oral syrup or

suspension.” [15]

15

Page 16: Modeling and simulation of photocatalytic degradation of ...

1. Motivationto lessen their effect on the environment.

In parts of the world with an intensive clothing industry (like China, Bangladesh, Viet

Nam, India, Turkey [20]), the attention is more frequently placed on dyes entering the

water cycle. Some dyes are toxic, mutagenic [21], or carcinogenic [22]. Therefore, the

removal of dyes from the wastewater is imperative. While the number of different

chemicals used in the dye industry is indeed small compared to pharmaceuticals, the

range of different substances is still extensive. In the European REACH3 database [23]

in the product category “Textile dyes, and impregnating products” (November 2018)

over 2150 chemicals were registered. While the overall production of around 900 thou-

sand tonnes of organic dyes in the year 2003, [24] is low compared to other chemical

substances, the amount that can be found in the environment is still significant. The

high volumes of water required during some dyeing processes inflate the amount of

produced waste drastically. In 1977 Anliker published a review titled “Color chemistry

and the environment,” where he stated from personal experience that around 10 -

20% of the active dyes are lost in the residual liquid waste [25]. Even though this re-

view was published more than 40 years ago, it is reasonable to assume that the order

of magnitude still applies for parts of the textile industry. In particular, in less devel-

oped countries outdated but low-cost production facilities are often still in use and

lead to high pollution levels in the clothing industry to this day.

Diverse efforts have been made to remove organic molecules from wastewater, such

as membrane filtration, activated carbon adsorption, and advanced oxidation pro-

cesses (AOP). AOPs are advantageous when the pollutants have a high chemical sta-

bility. Almost the total mineralization of contaminants to carbon dioxide, water, and

inorganic compounds can be achieved. If this is not the case, a partial oxidation can,

at least, help to make the initial pollutants more biodegradable [4, 26]. Different tech-

niques involve the generation of hydroxyl radicals, which are non-selective and have

twice the oxidizing power of chlorine [5, 7, 27, 28, 29]. The non-selectiveness is essen-

tial when dealing with such diverse pollutants as pharmaceuticals or dyes. Heteroge-

neous semiconductor photocatalysis has become an attractive method to remediate

environmental contamination due to its high photocatalytic activity, non-toxicity, and

photostability [4, 27, 30, 31, 32, 33]. Successful degradation experiments have been

reported [34] for various organic molecules, such as azo dyes [35], furfural [36], sul-

famethoxazole [37], and acetylene [31].

Understanding andmodeling the processes related to the degradation of organic mol-

3Registration, Evaluation, Authorisation, and Restriction of Chemicals

16

Page 17: Modeling and simulation of photocatalytic degradation of ...

ecules are essential for the development and characterization of new materials and

methods. This idea forms the core of this thesis. To establish a descriptivemodel, a set

of experimental systems is used as its foundation. These experiments are based on a

slurry mixture of photocatalytic nanoparticles to maximize the surface area of the sys-

tem. The up-scaling of such a setup can provide challenges regarding the separation

of the nanoparticles from the water after the treatment. As exemplary photocatalysts,

titaniumdioxide (TiO2), and zinc oxide (ZnO) were chosen and compared in their degra-

dation efficiencies. Even though several semiconductors have been studied for appli-

cations in wastewater decontamination, ZnO and TiO2 are frequently themost studied

photocatalysts because of their interesting optical properties, low cost, and availability

[38]. Although ZnO is usually described as the most active semiconductor [39], TiO2

is used more frequently because it is more stable than ZnO in aqueous solution [40].

Whenever photocatalytic systems are applied in an actual wastewater treatment plant,

a risk assessment regarding thematerial output into the environment is necessary due

to their photoactivity, size distribution, and potential toxicity for aquatic organisms in

the case of ZnO [41]. While such a study is out of the scope of this work, the properties

and risks of the used materials are discussed in more detail in Sec. 2.3 (p. 31).

The simulation of the kinetic processes, which are involved in the photocatalytic oxida-

tion of organic pollutants, are expected to lead to a deeper quantitative understanding

of the whole mineralization process. This understanding could help in optimizing the

design of reactors for photocatalytic water purification. A further issue of reactor de-

sign is the radiation transport in the reaction volume, investigated for example, in [42,

43, 44]. The degradation kinetics of organic pollutants has often been described within

the framework of the Langmuir-Hinshelwood (LH) model (see e.g., [45] and references

therein). For example, Minero [46] discussed several kinetic models of the photocat-

alytic process with the aim to obtain equations with physical meaning and reduced

complexity. Experimental validation of reported kinetic models has been performed

by Andreozzi et al. [47] for the photocatalytic degradation of 4-nitrophenol. The LH

model, combined with a Lambert-Beer type model for radiation transport, has been

applied to describe the degradation of methylene blue by Sannino et al. [48].

Generally, the photocatalytic mineralization of organic molecules is a complex process,

where parts of the initial molecules are oxidized step by step on the catalyst surface in

the presence of photo-generated radicals. These radicals can destroy the molecules’

bonds, thereby breaking them into smaller molecules. These molecules may then des-

orb from the surface of the photocatalyst. Consequently, various intermediate mol-

17

Page 18: Modeling and simulation of photocatalytic degradation of ...

1. Motivationecules of different sizes emerge in the solution. In this work, three advanced model

approaches of photocatalytic degradation are presented. These models extend the

single-species approach [45, 48, 49] by including the evolution of intermediates. In a

first case, the degradation by successive oxidation of the elements of the initial mole-

cules, excluding the possibility of molecule fragmentation, is studied. Within a second

complementary model, random fragmentation of organic molecules into smaller inter-

mediates and oxidation of the smallest fragments is considered. The last model aims

to correct the interpretation of the organic molecule as a simple chain through the

inclusion of excess bonds.

18

Page 19: Modeling and simulation of photocatalytic degradation of ...

2. Introduction

2.1. Modeling and simulation

Simulation in materials science aims to determine, understand, and predict the prop-

erties of actual materials. In the initial phase of computerized research, only simple

models could be evaluated. They often just described pure crystalline materials or ba-

sic reaction correlations. In the meantime, a large number of models are available and

deployable on increasingly powerful computational resources. Nowadays, it is possi-

ble to enhance research in a wide range of materials science topics with the aid of

simulation.

The creation of models is necessary independently of computer simulations, since it is

difficult to gain direct knowledge even in a clearly defined system. This predicament is

rooted in the problem that the monitored systems and their interactions cannot be in-

vestigated adequately or not directly in nearly all cases. For this reason, an abstraction

step is essential for the scientific process. The simple imitation of real relationships to

gain knowledge can be seen as the origin of modeling. The structures of a real system

are thereby reproduced in a similar but simplified way - a model.

For example, experiments can be interpreted as a particular form of modeling. In an

experiment, the challenge lies in translating the experiment setup into models and

back again. Without a clear abstraction layer, it would not be possible to generate

reliable and comparable data from experiments. In theoretical modeling, the primary

focus is more on changing the degree of abstraction. This evolution from elementary

models to complex systems to describe a given system better and better over time can

frequently be observed in science. Thus, constant efforts to describe our surrounding

19

Page 20: Modeling and simulation of photocatalytic degradation of ...

2. Introductionworld in more detail helped to understand it on new levels.

With the development of computer technology, increasingly sophisticated models of

real systems have become feasible. Old ideas to describe certain scientific aspects can

nowadays be used, due to the massive increase in computational power.

For example, Deep Learning is a topic that receives a tremendous amount of attentionin many areas of our daily and scientific life. The gross of the theoretical groundwork

was already completed in 1986 in work presented by Rumelhart, Hinton, and Williams

on convolutional neural networks [50]. However, it took over 25 years before the the-

oretical found methods would be used on the wide-scale we see today. Besides the

growth in processing power, additions to the initial ideas played a critical role as well.

Scientists were long fascinated with the idea to describe the behavior of atoms and

molecules withmathematical rules. One of themethods to describe these interactions

are molecular dynamics simulations, which are already conducted since the 1950s. In

this field, the growth of computational power can be very impressively demonstrated

by the systems that are investigated. Alder and Wainwright showed in their 1959 pub-

lication [51] experiments with 32 to 108 particles but claimed already that ‘Computers

now being planned should be able to handle ten thousand molecules in calculations

which do not require very many collisions.’ With the development of better force field

descriptions (Lennard-Jones potential) andmore robust computational systems, more

realistic systems could be tackled. So could Rahman calculate a good approximation

for self-diffusion of Argon in 1964 [52]. Roughly ten years later, in 1975, the first suc-

cessful folding of a protein was published by Levitt and Warshel [53]. The addition of

watermolecules as solvents was possible around 1988 to describe the behavior of pro-

teins more reliable [54]. Another critical dimension that developed was the length of

the time axis that could be explored. Duan and Kollman were able to observe protein

folding in water for 1µs instead of the 200ps ten years prior [55]. A very impressive

milestonewas themodeling of the Satellite TobaccoMosaic Viruswith up to 1,068millionatoms in 2006 by Freddolino et al. [56].

Nowadays, it is possible to create a virtual system in the rough likeness of a real sys-

tem. As symbolically show in Fig. 2.1, all three pillars of modern science: (i) theory, (ii)

experiment, and (iii) simulation are part of the overall process. In a traditional setup,

just the left half in this representation (theory and experiment) would be used. A the-

ory would be created to describe a system, and a loop of observing and manipulating

the real system would generate data to bring the theory to a descriptive model. A

20

Page 21: Modeling and simulation of photocatalytic degradation of ...

2.1. Modeling and simulation

Model

Theory Experiment Simulation

SystemReal Virtual

System

abstract

constrain

create

observe

parameterize

obser

ve

initialize

manipulate

manipulate

Figure 2.1.: Relationship between the model and the real and virtual systems.

virtual system can now be created by simulating the model. The boundaries and con-

straints of the new systems are mostly derived from the initial theory-crafting process.

The experimental observations generate initial values and properties. After the model

is parameterized and prepared for the simulation, observations can be made. The

observation manipulation loop is much faster and easier to drive compared to the

experimental loop. This tight loop is one of the main advantages of using a virtual

system.

For a long time, computational simulations were just used to confirm experimental

findings or to interpret them. In these cases, the model is already established and

often well examined by experiments. With the higher level of available computational

power, simulations can cover vast parameter spaces while running the manipulation /

observation loop. This extension opens the option to start with just a few or without

experimental data at all. A constant reiteration of the model based on the received re-

sults is often implemented. The virtual system in such an instance can be used to pre-

dict the outcome of experiments in the real world. Especially for very time-consuming

experiments, this method can be used as a powerful tool to preselect promising candi-

dates. It is important to note that the results of a virtual system without experimental

verification should always be treated with caution.

The goal often seems to be the perfect model. Here it must be noted that someone

21

Page 22: Modeling and simulation of photocatalytic degradation of ...

2. Introductionwhowould be able to create such a perfectmodel of a systemwould, in fact, not need it.

This person could directly understand the original in its whole complexity. Especially in

materials science, wheremanymechanisms on different sizes and time scales interact,

a single monolithically approach is not likely to be successful. Often individual models

have to be cross-linked to obtain application-relevant information. In this work, too, the

overall picture of the system is created by combining different methods on different

scales to obtain a better understanding of the examined system.

2.2. Heterogeneous photocatalysis

2.2.1. History

The influence of light on our environment is a topic that is connected with humankind

for eons. Chemical reactions that were driven by light were first discovered around

1790 by Joseph Priestley (1733-1800). His experiments with nitric acid can be seen as

the start of modern photochemistry [57].

The interest in catalytic photoreactions started in the 1900s. An example is the oxida-

tion of oxalic acid to formic acid with uranyl as catalytic material in 1911 by Bruner and

Kozak [58]. Even with their research limited to sunny spring and summer days, they

could already establish many essential connections like between the reaction rate and

the light intensity or the higher efficiency of blue/violet light. The word photocatalysis

was often used in the beginning 1900s to describe photosynthesis [59, 60]. This kind

of reactions would not be counted to the photocatalysis definition nowadays. The

International Union of Pure and Applied Chemistry (IUPAC) [61] defines the term pho-

tocatalysis as:

“Change in the rate of a chemical reaction or its initiation under the action

of ultraviolet, visible, or infrared radiation in the presence of a substance

—the photocatalyst— that absorbs light and is involved in the chemical

transformation of the reaction partners.”

and photocatalyst accordingly:

“Substance able to produce, by absorption of ultraviolet, visible, or infrared

radiation, chemical transformations of the reaction partners, repeatedly

22

Page 23: Modeling and simulation of photocatalytic degradation of ...

2.2. Heterogeneous photocatalysiscoming with them into intermediate chemical interactions and regenerat-

ing its chemical composition after each cycle of such interactions.”

In the modern interpretation, heterogeneous photocatalysts are often semiconduc-

tors. First, mainly bulk semiconductors systems were studied to understand the sur-

face properties and reactions under visible irradiation. The area did not see much

activity due to problems with long-term stability and high cost. A new spark was cre-

ated with the now exceptionally highly cited Nature paper by Fujishima and Honda in

1972 [62]. In this publication, they describe the use of TiO2 to split water. A further

accelerator was the first oil crisis in 1973. During this time, the over-dependence on

fossil fuel got apparent, and the possibility to create fuel through water splitting from

light was encouraging. It was evident early on to the scientist in this field that using

light to foster chemical reactions will be an essential part of creating a sustainable fu-

ture. Already in 1912, Giacomo Ciamician stated in his New York speech about the

Photochemistry of the Future [63]:

“So far, human civilization has made use almost exclusively of fossil solar

energy. Would it not be advantageous to make better use of radiant en-

ergy?”

However, the first iterations based on the findings of Fujishima and Honda still battled

with problems in long-term stability and efficiency. Intertwining of nanotechnology in

the field during the last decades helped to overcome some of the downfalls. Nanos-

tructured photocatalysts are now the most promising pathway. In this section, the

physical and chemical properties of the bulk semiconductor and the critical effects of

the nanostructure are discussed.

2.2.2. Semiconductor band structure

A more in-depth look into the electron structure of the explored semiconductors is

necessary to approach the process of photocatalysis. In a commonly used depiction,

the electron structure is described by bands of allowed energy levels. These bands

are ordered from lowest to highest energy. The last band that is filled with electrons

in the ground state is called the valence band (VB). Energetically directly above, as the

first empty band, is the conduction band (CB). The distinguishing feature of a semicon-

ductor band structure is the noticeable gap between VB and CB.

23

Page 24: Modeling and simulation of photocatalytic degradation of ...

2. Introduction

EvacEcEFEv

Egap

Figure 2.2.: Energy diagram for a typical n-type semiconductor. Showing the energy

levels for vacuum Evac, conduction band EC and valenace band EV, as wellas the Fermi engery EF and the band gap Egap.

In Fig. 2.2 a simple energy diagram for a typical semiconductor is shown. The filled

states are depicted in blue and are confined by the energy level of the valence band

EV. The empty bands between the energy level of the conduction band EC and the

vacuum level Evac are shown in red. The vacuum level describes the energy that is

needed to remove an electron from the system. Both are isolated by the bandgap

Egap. The Fermi level EF describes the energy level, which would have a 50% chance to

be occupied by an electron. It is located inside the bandgap.

The bandgap is a fundamental property to understand the photocatalytic process.

Photons can interact and excite electrons in the valence band and lift them over the

gap into the conduction band, leaving a hole in its place. The created electron-hole

pair, also called an exciton, is the source of the photocatalytic activity. In addition to

the energy difference separating the valence and conduction band, the gap type is

essential. To investigate the different kinds of bandgaps, the simple representation as

used in Fig. 2.2 is not sufficient. Describing the bands in more detail is not feasible in

normal space. Due to Heisenberg’s uncertainty principle, the exact energy and posi-

tion cannot be determined at the same time. With the help of the reciprocal space,

the problem can be transformed in such a way that it enables us to describe the mo-

mentum and energy at the same time. The wave vector k can be seen as a positionalvector in the reciprocal space. Based on the essential points in the first Brillouin zone,

the different possible energy states can be plotted as bands. In Fig. 2.3 band struc-

tures for two different phases of TiO2 are shown as examples. It is clear visible, that

the energy level of the valence and conduction band change with k. In the first graph

24

Page 25: Modeling and simulation of photocatalytic degradation of ...

2.2. Heterogeneous photocatalysisshowing anatase, the maximum of the valence band (red dot) is not in the same posi-

tion as the minimum of the conduction band (blue dot). The bandgap is indirect. An

excited electron will need to change its momentum alongside its energy level. This

means that besides the interaction with the photon, the electron has furthermore to

interact with a vibration of the lattice (phonon interaction). This additional necessary

step lowers the efficiency of these kinds of materials. In contrast, in rutile (Fig. 2.3 b)

a direct bandgap can be found at the center of the Brillouin zone (Γ). A photon can

efficiently excite an electron directly to overcome the gap [64].

The presented band structures were calculated for the AFLOW4 repository [65] using

the density functional theory. Compared to experimental results, these calculations

underestimate the bandgap significantly. For the rutile phase, for example, the experi-

ment measures a bandgap of around 3.0 eV [66, 67], and the here shown calculations

estimates just 2.3 eV.

-4

-2

0

2

4

Γ X Y S Γ Z S1 N P Y1Z|X P

a)

Wave vector k

Energy

E–E F(eV)

Γ X M Γ Z R A Z|X R|M A

b)

TiO2 anatase TiO2 rutil

Figure 2.3.: Band structure of TiO2 in the a) tetragonal anatase phase and b) tetragonal

rutile phase, created from data provided by the AFLOW repository. The

red dots mark the lowest unoccupied state, while the blue dots mark the

highest occupied state.

4aflow.org – Automatic FLOW for Materials Discovery

25

Page 26: Modeling and simulation of photocatalytic degradation of ...

2. Introduction

2.2.3. Interface between a semiconductor and a redox electrolyte

In the dark

The Fermi level describes an electrochemical potential for electrons. At absolute zero,

all energy levels below the Fermi level are occupied. The distribution of occupied /

empty states around the Fermi level is described in the Fermi–Dirac statistics.

FF(εe) = 1

exp

( εe–EFkBT)+ 1

(2.1)

If the energy of a band ismuch larger than the Fermi level Equation 2.1 is approximately

zero. The band is empty. In the case of much lower energy, the distribution converges

to one, and therefore these bands are filled. As mentioned above, if the energy level

of a band is the same as the Fermi level, it has a 50% chance to be occupied by an

electron. In a perfect semiconductor, the Fermi level would be in the center of the

bandgap, but impurities or doping shift this position. In n-type semiconductors, the

additional electrons in the conduction bandmove the Fermi level towards it. For p-type

semiconductors, the Fermi level is closer to the valence band. Further, the different

curvature of the valence and conduction band can influence the position of the Fermi

level.

The semiconductor is not isolated during the photocatalysis process. In our case, the

exciting aspect is the behavior of the material in conjunction with a solution containing

a pollutant. Between the redox pair and the semiconductor, an electronic equilibrium

will be established, as soon as they come into contact. The process is quite similar to

the combination of semiconductors and metals or differently doped semiconductors.

The reaction can be abstracted as a simple electron transfer between an oxidized (Ox)

and a reduced (Red) species.

Ox + z e–

Red {1}

Written in terms of electrochemical potential:

26

Page 27: Modeling and simulation of photocatalytic degradation of ...

2.2. Heterogeneous photocatalysis

μOx + μe = μRed (2.2)

The Nernst equation is now used to describe the redox potential.

U = U◦–RTzF ln

CRedCOx (2.3)

U = U◦–kBTze ln

CRedCOx (2.4)

The redox fermi level can be defined relative to the vacuum level [68] (as recommended

by the IUPAC as a value of −4.4 eV is used [69]). With the standard reduction potential

U◦redox

based on the standard hydrogen scale we can write:

EF,redox = –4.4 eV – eUredox (2.5)

Uredox = U◦redox

–kBTze ln

CRedCOx (2.6)

EF ,redox = –4.4 eV – eU◦redox

+kBTz ln

CRedCOx (2.7)

As stated before, the Fermi level can be interpreted as the electrochemical potential of

the electrons. Because we investigate the system in an equilibrium state, the potential

of the electrons in the semiconductor and the redox system is equalized. When the

n-type example semiconductor would come into contact with a redox-pair with a lower

Fermi level, some electrons near the interface will switch over to the oxidized species in

the solution. This will lead to a depletion layer in the semiconductor and hence to the

bending of the bands near the interface. The resulting band structure at the interface

is very similar to a Schottky barrier that would be formed by a semiconductor/metal

interface [70].

27

Page 28: Modeling and simulation of photocatalytic degradation of ...

2. Introduction

Evac

EcEF

EvEgap

Ebend

electrolyte

νh

δp0 Wdl Wdl + Lp

electrolyte

in dark under illumination

Figure 2.4.: Energy diagram for a typical n-type semiconductor in contact with an elec-

trolyte in dark and under illumination.

Under illumination

If a semiconductor is illuminated, photons can excite electrodes in the valence band.

The energy of the photon is proportional to its frequency ν. When enough energy

hν > Eg is transferred to the electron, it can jump over the bandgap into the conductionband. This leaves a hole in the valence band that behaves like mobile positive charge

carriers. The absorption coefficient α in the case of a direct bandgap is listed in Eq. 2.8.For an indirect bandgap Eq. 2.9 is valid. In these two equations A′ is a proportionality

constant. This relationship can inn addition be used to determine optical bandgaps

using a Tauc plot [71, 72].

α = A′(hν – Eg)0.5 (2.8)

α = A′(hν – Eg)2 (2.9)

Based on the definition of the absorption coefficient, we can make statements about

the penetration depths δp. The definition is based on the reduction of the initial irra-diation flux I0 by a factor of Euler’s number e. A variant of the Lambert-Beer law gives

the relation between the flux strength and the position in the material.

28

Page 29: Modeling and simulation of photocatalytic degradation of ...

2.2. Heterogeneous photocatalysis

I(x) = I0 exp (–αx) (2.10)

Combined with the definition of the penetration depths, a simple relation can be es-

tablished.

I(δp) = 1

eI0 (2.11)

1

eI0 = I0 exp (–αδp) (2.12)

δp = 1

α (2.13)

From this relationship, we can conclude that penetration depths can be vastly differ-

ent depending on the type of bandgap that has to be overcome. A typical depth for

semiconductors with a direct bandgap is in the range of 0.1 to 1 µm. For materials

with indirect bandgaps, the penetration depths can be longer by order of magnitudes.

An example is gallium phosphide (GaP) at a wavelength of 515nm with a penetration

depths of 10µm [73].

For the practical usage of photolysis, it is crucial to investigate the fate of the created

excitons. The first question is how many of the new charge carriers are available to an

external system or circuit. The proportion between all created and collected carriers

are described by the quantum yield Φ. This critical parameter in the photochemistryis still a significant and not fully solved challenge to measure and to model. Alone

measuring the light that reaches the semiconductor is nearly impossible due to the

many different material layers in an experimental setup that can absorb or scatter the

light by themselves.

Despite the problems in modeling the quantum yield, it is possible to do an upper

limit estimation. Most important is the interactions near the interface between the

semiconductor and the electrolyte. In the exemplary n-type semiconductor, some of

the electrons are transferred into the electrolyte. Due to this shift, the band is bent

up toward the surface, as discussed in the dark section before. The width of this layer

that is depleted by the majority charge carriers is described by Eq. 2.14 based on the

density of donor/acceptor states Ns the relative permittivity of the semiconductor ε

29

Page 30: Modeling and simulation of photocatalytic degradation of ...

2. Introductionand the potential difference by the bent bands Ebend [74].

Wdl = √2εε0EbendeNs (2.14)

The electric field separates Electron-hole pairs that are created in the depletion layer.

The minority charge carrier is moved towards the interface. In our example, the holes

are guided out, and the electrons are moving in the opposite direction, into the bulk

of the semiconductor. So the holes may react with the redox system in the electrolyte.

Pairs that are created away from the depletion layer are bound to recombination.

There is a chance that charge carriers still reach the depletion layer and, therefore,

the interface by diffusion before the recombination occurs. This depends on the life-

time τ of theminority charge carrier and theirmobility μ. In Eq. 2.15 the diffusion lengthfor holes Lp is defined [75].

Lp = √Dpτp (2.15)

Lp = √kBTμpτp (2.16)

These two dimensions must now be put in relation to the penetration depths δp. In Fig.2.4 the relative positions of all these lengths are presented. Because this is an upper

limit estimation, we can use the assumption that all separations that happen between

the surface and Wdl + Lp contribute to the carrier flow to the interface. The resulting

relation (Eq. 2.17) is the Gärtner equation [76].

Φ = 1 –

exp

(–Wdlδp)

1 +Lpδp

(2.17)

Φ = 1 –exp (–αWdl)1 + αLp (2.18)

This equation has many drawbacks that limit it as an estimation of the system. For

30

Page 31: Modeling and simulation of photocatalytic degradation of ...

2.3. Photocatalytic materialexample, it is just valid if the electron transfer on the interface is fast. If the reaction is

slow, a build-up of the minority charge carriers will start to change the band structure.

A result of these flatted bands would be a significant increase in recombination in the

charged depletion layer that undercuts the assumption made in the beginning [70].

2.3. Photocatalytic material

2.3.1. Overview

In general, semiconductors are good candidates to exhibit photolytic behavior. As

explained in the section before, the bandgap that separates the valance and conduc-

tance band is a significant value. The wider the gap is, the more energy a photon

needs to create an electron-hole pair in the material. Bandgaps over 3.0 eV, for ex-

ample, lead to a photon absorption edge in the near ultraviolet (UV) region. Figures

2.5 - 2.7 shows the wide range of possible wavelengths. The illustration is based on

Aracely Hernández-Ramírez and Iliana Medina-Ramírez [77] extensive compilation of

photocatalytic material properties.

The most commonly used materials can be found under the metal oxides. TiO2 and

ZnO are prominent members of this group and from particular interest in this thesis.

As such, they will be described in more detail in separate sub-sections following this

overview.

One interesting example of this group on the high end of the energy spectrum (5.0 eV)

is zirconium dioxide (ZrO2). As a stable and already widely used oxide in ceramic tech-

nology, it is very well studied. Based on the high bandgap, the produced electron-hole

pair has a high chemical potential. Besides, the bandgap can be modified significantly

through the preparation of the material [78]. Compared to TiO2 the photolytic activity

of ZrO2 is quite low [79]. Another drawback is the absorption edge at 248nm, which

means solar irradiation cannot be used with this semiconductor.

The second example is tungsten trioxide (WO3), in which the absorption edge is further

shifted into the visible spectrum compared to TiO2 and ZnO. With a bandgap of 2.8 eV

up to 8.5%of the solar spectrum could be utilized. Besides, the possible use of sunlight

WO3 is equally interesting because it is very stable under many conditions, like strongly

acid solutions [80]. The photocatalytic efficiency of pure WO3 is limited due to the

31

Page 32: Modeling and simulation of photocatalytic degradation of ...

2. Introduction

ZrO2

Ga2O3

SnO2

Cr2O3

Nb2O5

Nd2O5

TiO2

Ta2O5

ZnO

CeO2

Bi2O

3

In2O3

WO3

Fe2O3

V2O5

Ag2O

1.21.31.51.72.02.53.04.0E (eV)300 500 600 800 1000400 900700λ (nm)

Figure 2.5.: Bandgap energies and absorption edges for selected metal oxides [77].

conduction band position relative to the oxygen / oxygen-radical pair. Therefore, the

electrons cannot be scavenged easily from the conduction band, and electron-hole

pairs are more likely to recombine. With a co-catalyst like noble metals, this problem

can be attenuated [81].

The last example for metal oxides in Fig. 2.5 is silver oxide (AgO2). This material could

theoretically use up to 76.3% of solar irradiation. Even with the low energy of the

exciton that would limit the number of organic compounds that could be degraded,

a photocatalytic material that could use the full visible spectrum would be beneficial.

In practical scenarios, however, high photosensitivity causes serious problems. So

is the oxide not stable under light and decomposed into silver and oxygen. There

are approaches to stabilize the material for the use as photocatalyst [82], but broad

adoption is not to be expected due to the still low efficiency and high cost.

The photocatalytic materials should be used in wastewater cleaning. It is essential

to consider the environmental impact of the photocatalytic materials in. Besides, a

low photostability, silver oxide is further highly toxic for aquatic ecosystems. Other

materials like gallium(III) oxide (Ga2O3), zinc oxide (ZnO), cerium(IV) oxide (CeO2) or

vanadium(V) oxide (V2O5) share the same weakness in different level of severeness.

In the realm of chalcogenide (Fig. 2.6), the most investigated materials are cadmium

and zinc sulfide. Both are promising for the production of hydrogen due to the highly

negative potential of the excited electrons. The main downside with most sulfide com-

pounds is their instability in solution under irradiation, which is as well the reason for

the high toxicity of the material for aquatic systems. Sacrificial reagents can be added

32

Page 33: Modeling and simulation of photocatalytic degradation of ...

2.3. Photocatalytic material

GaP

β-Ge3N4

SiCTa3N5

GaAs

ZnS

Sb2 S5

CdS

MoS2

CdSe

CdTe

Bi2 S5

1.21.31.51.72.02.53.04.0E (eV)300 500 600 800 1000400 900700λ (nm)

Figure 2.6.: Bandgap energies and absorption edges for selected chalcogenides (•) andother binary compounds (N) [77].

to the system, to lessen the anodic photocorrosion. Through different approaches,

the systems can be stabilized enough to be possible candidates to produce chemical

energy directly from the sunlight. This is even the case without a solution for long-term

stability since the material is mostly considered an unwanted byproduct in oil refining.

Thereby generating hydrogen, waste is eliminated. Even with some experiments using

cadmium or zinc sulfide to degraded pollutants, the materials are not considered in

the setting of this work due to their stability and toxicity problems [83, 84, 85].

1.21.31.51.72.02.53.04.0E (eV)300 500 600 800 1000400 900700λ (nm)

Figure 2.7.: Bandgap energies and absorption edges for selected ternary (•) and qua-ternary semiconductors (N) [77].

An astonishing number of semiconducting materials were tested in the last decades

for their photocatalytic properties. Ternary and quaternary semiconductors like met-

allates, oxysulfides, oxyhalides, oxynitrides, or oxysulfide represent a large portion of

all potential photocatalytic material (Fig. 2.7). Some of these materials show promising

properties for water splitting. However, for use in water detoxification, none of the

materials can rival the properties of TiO2 if UV radiation is considered. In the reviews

by Hernández-Alonso et al. [86] and Di Paola et al. [87] the properties of the further

photocatalytic materials is collected. A recollection of this information here would go

beyond the scope of this work.

33

Page 34: Modeling and simulation of photocatalytic degradation of ...

2. Introduction

2.3.2. Titanium dioxide

Since the publication by Fujishima andHonda in 1972 [62], titaniumdioxide TiO2 repre-

sents still the highest echelon for photocatalyticmaterial. Any newmaterial is inevitable

compared to this standard. In this work, TiO2 will as well serve as the foundation.

Titaniumdioxide is a solid that ismostly prepared as a powder with a very brilliant white

color due to the very high refraction index of 2.6. It is odor- and tasteless. The oxide

is very stable in a wide variety of environments and unsolvable in water. Commonly

the material consists of a mixture of different crystalline types. Although TiO2 can be

forced into multiple different phases, two are of significant interest.

Unit cell:

c = 2.9587 Å

RutileSpacegroup: P42/mnm

a = 4.5937 Å

a)

Unit cell:

c = 9.5143 Å

AnataseSpacegroup: I41/amd

a = 3.7845 Å

b)

Figure 2.8.: Crystal structure of (a) rutile (8 unit cells) and (b) anatase (4 unit cells) from

different perspectives.

First rutile, the most commonly occurring natural form of titanium dioxide. The cristal

structure is tetragonal (P42/mnm) as shown in Fig. 2.8 a. Two atoms form the unit cell

of this structure together. Every titanium atom is surrounded by six oxygens forming

connected octahedrons. Each octahedron is in contact with ten others (two edge and

eight corner connections). It is worth noting that the octahedrons are slightly distorted.

Based on this cristal system the bandgap has a value of around 3.0 eV (see Fig. 2.3 a)

[88]. The density of rutile amounts to 4.13 g cm−3 [89].

Second anatase, the crystal phase that is commonly linked to a higher photocatalytic ac-

34

Page 35: Modeling and simulation of photocatalytic degradation of ...

2.3. Photocatalytic materialtivity. The crystal structure is tetragonal (I41/amd) as shown in Fig. 2.8 b similar to rutile.In contrast to rutile, four atoms are located in the unit cell, and the octahedrons are

stronger distorted. While the octahedron structure stays the same, the interconnec-

tion changed. So are in anatase each octahedron in contact with just eight neighbors

(four edge and four corner connections). This results in altered distances between

Ti-Ti (extended) and Ti-O (shorten). Overall the density is reduced by 0.37 g cm−3 to

3.76 g cm−3 compared to rutile [89]. The bandgap is also slightly elevated to 3.2 eV.

While the valence band edge is very comparable between both materials is the con-

ductance band is roughly 0.2 eV higher in anatase [67].

The thermodynamic stabilities of the different titanium dioxide phases are quite sim-

ilar. Due to the small disparities, the stability order for macrocrystalline systems is

rutile → brookite → anatase. While annealing at temperatures below 400 °C will yield

anatase at higher temperatures (over ~600 °C), rutile becomes the dominant phase.

In the case of nanomaterial, surface energy plays an important role. For tiny particles

under 11nm, anatase is themost stable configuration. If the particle size is over 35nm,

rutile is again the most stable phase. In between, the brookite is most favorable [90,

91]. While it is still heavily debated if anatase or rutile is the superior photocatalytic

material, it seems very clear that a mixture of the two phases shows highly beneficial

properties [92]. The practical standard for photocatalytic materials Degussa’s P25 is

likewise a mixed-phase product. It is compiled form 75% anatase and 25% rutile. The

primary challenge in understanding the performance of P25 is to characterizing the

powder. So far the precise distribution of the phases and the surface futures is still

not fully understood. The general explanation regarding the enhanced activity refers

to the interfaces between both phases as the cause of the cooperative effect. This is

one of the reasons, while ultrasonication of P25 shows just a very limited effect [93].

While the accessible surface area is highly increased, and the efficiency of photon ab-

sorbance is improved by separating the particle agglomerations, the overall reaction

rate is just raised slightly. Possible explanations of the effect include the formation

of heterojunctions, due to the different band configuration. Besides, charge transfers

between the phases through band bending is likely [88].

Using TiO2 brings the significant benefits of a very well established material. Through

its high usage, the interaction with the environment and humans is well investigated. In

2015 the worldwide estimated consumption was 5.89 million tonnes, but just around

10 thousand tonnes are attributed to nanomaterials [94]. Primary use cases can be

found in the paint and coatings industry (55%), as filler in polymers (26%) and paper

35

Page 36: Modeling and simulation of photocatalytic degradation of ...

2. Introductionproduction (8%). In the remaining 11% are mostly applications in the textile industry,

ink production, food preparation, cosmetic formulation, and as a catalyst [95]. Primar-

ily through the use in the cosmetic and food industry is titanium dioxide well tested. As

food additive (E171), it is categorized in the European Union in “Group II: Food colours

authorised at quantum satis5”. As food color, it shares this group with a few other sub-

stances like Beetroot Red (E162) or caramel (E150). It can be found in nearly all daily

foods that need white color or to be brightened from prominent products like tablets,

chewing gum, cheese, ice cream, noodles over flavored drinks, and cocktails to salads

and fine bakery wares [96]. As a result of an intensive discussion about nanomateri-

als in cosmetic products around the year 2007 in the EU, an extensive safety report

about nano-scale titanium dioxide in suncream was compiled over the next years [97].

It is important to note that the investigated studies in this report mostly concerned

coated material. Furthermore, TiO2 powder produced for UV protection in cosmetic

has been optimized to show as little photocatalytic activity as possible. Therefore, it

is still necessary to ensure that any application with titanium dioxide releases as little

material as possible into the environment.

2.3.3. Zinc oxide

The high similarity with TiO2 brought ZnO quickly in the spotlight of the photocatalytic

research [98]. Even with just 41% of the recent publication count compared to tita-

nium dioxide6, it is still the second most published about photocatalyst.

Zinc oxide is a white, odorless n-type semiconductor that is commonly available as

a powder, but single crystals can be produced as well. It is practically unsolvable in

water. Growing ZnO on a cubic lattice can produce a zincblende structure, but this

configuration does not play an essential role in photocatalysis. The most important

crystal structure is wurtzite, that is formed at ambient condition. As shown in Fig. 2.9

the crystal is hexagonal (P63/mc). Each zinc is surrounded by four oxygen. This tetra-hedral coordination is also accurate from the perspective of an oxygen atom. The

direct bandgap is similar to TiO2 with 3.37 eV [99]. ZnO needs, therefore, near UV ir-

radiation to show photocatalytic activity. Due to the non-centrosymmetric structure,

piezoelectric properties can be observed in zinc oxide [100].

The global production of zinc oxide was estimated at 1.37 million tones for 2014 [101].

5as much as is sufficient6Indexed in Scopus for the year 2017

36

Page 37: Modeling and simulation of photocatalytic degradation of ...

2.3. Photocatalytic material

Unit cell:

c = 5.307 Å

ZnO WurtziteSpacegroup: P63/mc

a = 3.289 Å

Figure 2.9.: Crystal structure of zinc oxide - wurtzite (9 unit cells) from different per-

spectives.

Nano-sized materials account, however, just for several thousand tones [94].

Its impact on the environment limits the use of ZnO to degraded organic components

in wastewater. In the literature, many different perspectives on the topic are available.

In a review in 2016 Lee et al. [99] stated about ZnO:

“ZnO is an environmental friendly material as it is compatible with living

organisms, which lending itself nicely to a broad range of daily applications

that will not leave any risks to human health, and environmental impacts”

This is a strong claim, that is disappointingly not corroborated with references. In an

eco-toxicity comparision between TiO2, SiO2 and ZnO by Adam et al. a different picture

emerges [41]. There, it is clear that ZnO has the strongest influence on the growth

of the studied test organisms (Gram-positive Bacillus subtilis and Gram-negative Es-cherichia coli). Furthermore, the classification of zinc oxide as hazardous to water (WGK2: wassergefährdend) in Germany shows that the usage of ZnO requires special care,

to reduce the chance of environmental contamination [102]. In the European Union,

ZnO is classified as hazardous to aquatic systems (Aquatic Chronic 1 | H410: Very toxic

to aquatic life with long lasting effects) as well. While many studies show the effect of

zinc oxide on systems like microorganism, algae, or zebrafish embryos, the overall

ecotoxicological knowledge is still not sufficient to extrapolate to other circumstances

[103].

37

Page 38: Modeling and simulation of photocatalytic degradation of ...

2. IntroductionA further drawback of ZnO is the existence of anodic photocorrosion, which is espe-

cially problematic at lower pH values. The reaction sequence is listed below [104].

ZnObulk + 4 νh ZnObulk + 4e–+ 4h

+{2}

O2–s + h

+O

–s {3}

O–s + O

2–s + h

+(O O)

2–{4}

(O O)2–+ 2h

+O2 {5}

2 ZnO + 4 νh 2Zn2+(aq) + O2 + 4e

–{6}

During the process, the surface oxygen Os is detached from the bulk oxide. This pro-

cess reduces the catalytic activity of ZnO over time. Besides, Zn2+ ions will be released

into the environment at higher levels than possible by usual dissociation. This release

is especially problematic for the environment because Zn2+is allegedly the main rea-

son for the ascertained toxicity of ZnO. Different strategies are available to reduce the

effect. For example, passivation of the surface with polyaniline reduced graphene ox-

ide or fullerenes. Optimization of the exposed crystallographic surface influences the

strength of the effect [99].

Zinc oxide can be produced in a wide variety of morphologies [105]. The availability of

specific nanostructured surfaces is very promising for use in complex photoreactors.

The generation of column-to-rod morphological transitions, oriented nanocolumns,

and nanoplates can help to optimize surface area and to reduce photocorrosion.

Zinc oxide often presents a higher photocatalytic activity compared to titaniumdioxide.

The problems regarding photo-corrosion and solubility in strong acids and alkalismake

it challenging to including ZnO in a water cleaning system. All in all is zinc oxide is a

good model system in tandem with titanium dioxide.

38

Page 39: Modeling and simulation of photocatalytic degradation of ...

2.4. Light sources

2.4. Light sources

2.4.1. Solar

Solar irradiation is a universal source of energy available everywhere on earth in vary-

ing intensities. The idea of utilizing this source as efficiently as possible is gaining now

more andmore popularity. As stated in the historical introduction into heterogeneous

photocatalysis, it was already realized at the beginning of the 20th century that solar

irradiation is a promising energy source for chemical processes [63]. Nowadays, this

route in chemical synthesis is rediscovered in combination with the use of photocat-

alytic material to enable new pathways [106].

For the cleaning of wastewater, the usage of solar irradiation enables unique advan-

tages. First, the cleaning setup can be designed without the need for a highly reliable

electronic power-grid. Second, no additional burden on the environment through the

use of toxic materials common in some lamps like mercury. The final advantage is the

low running cost.

The most common photocatalytic materials can use a small portion of the solar spec-

trum. Titanium dioxide, for example, can utilize between 4.5% and 7.5% of the avail-

able solar irradiation (depending on the bandgap) [107]. Due to different drawbacks,

photocatalysts with a naturally smaller bandgap are rarely used. The main focus in

visible light photocatalysis is placed on the improvement of the established materials.

Many different strategies to treat TiO2 are available [108, 109, 110].

One approach is the combination with other photosensitive semiconductors. Two

main prerequisites are necessary for such a combination. First, the light absorption

edge of the added semiconductor needs to be well in the visible range. Second, the

level of the conduction band of TiO2 needs to be lower than the one of the added sen-

sitizer. This is important to enable an excited electron to pass easily from the added

semiconductor onto the titanium dioxide. A system formed from nanosized TiO2 and

cadmium sulfide (CdS) exhibit such a behavior [111]. There the photo-generated elec-

trons generated in the sensitizer (CdS) can be efficiently transferred to the conduction

band of titanium dioxide. This electron transfer enables an effective decomposition of

contaminants under visible light.

Doping is a useful tool to alter the properties of semiconductors. With the addition of

39

Page 40: Modeling and simulation of photocatalytic degradation of ...

2. Introductionmetal ions, new states can be introduced inside the bandgap. Often transition metals

are used since their redox energy states lay in between the bandgap of TiO2. The new

states introduced close to the valence or, respectively, the conductance band leads to

a redshift and therefore enables visible light absorption. Dopant concentrations need

to be optimized carefully, as quickly unwanted effects can overshadow the achieved

improvement. Mainly the introduction of new recombination sites causes problems,

reducing the average lifespan of the excitons, and therefore reducing the maximum

length they can diffuse. Electron-hole pairs created further away from the surface are

lost for the reduction of pollutants.

Non-metal ions can be beneficial dopants for TiO2 as well. While there is work on us-

ing a wide range of elements like fluorine, iodine, or phosphor as dopants, the primary

attention is focused on nitrogen and carbon. Sato discovered in 1986 [112] that the

addition of ammonium hydroxide (NH4OH) into the preparation process of TiO2 en-

ables a respond under visible light. In 2001, Asahi et al. [113] restarted the interest in

nitrogen as a dopant. It could subsequently be shown that an N-doped TiO2 material

can achieve an up to nine-time better degradation of methylene blue compared to

P25 in visible light (>450nm) [114].

One way to introduced carbon into titanium oxide is during the preparation via flame

pyrolysis [115]. Depending on the used fuels (glycine, hexamethylenetetramine, or ox-

alyldihydrazide), a significant change of the bandgap can be reached. With glycerin as

fuel, the absorption edge could be pushed to 435nm with a slow reduction of adsor-

bents until 560nm. This redshift covers a more substantial part of the solar spectrum,

compared to pure titanium dioxide. While the activation in the visible light is favorable,

it needs to be noted that the oxidation power and mobility of the generated holes

are lower than in pure TiO2. The combination of p- and n-dopant like nitrogen with

fluorine is an opportunity to stabilize the material system and reduce the increased

exciton recombination typical for doped materials. This co-doped systems still work

efficiently under visible light [116].

Besides incorporating elements directly into the structure of titanium dioxide, decora-

tion of the surface can have a considerable influence on the photocatalytic properties.

The introduction of noble metal clusters on the surface enhances foremost the life-

time of excitons, as they act as electron traps. They can be beneficial in shifting the

light absorption edge aswell. Particular experiments with silver showpromising results

[117]. One possible explanation presented in the published work is surface plasmon

absorption from the silver cluster.

40

Page 41: Modeling and simulation of photocatalytic degradation of ...

2.4. Light sources

irradiation in W/m2

125

250

150

175

200

225

100

500 km

Direct horizontal

Figure 2.10.: Long-term average of direct horizontal irradiation (solar radiation re-

ceived from above by a surface horizontal to the ground) for central Eu-

rope (top) and south-east Asia (bottom). Solar resource data obtained

under "Creative Commons License Attribution 3.0 IGO" from the Global

Solar Atlas, owned by the World Bank Group and provided by Solargis.

The design process needs to be optimized for it from the start to use solar energy effi-

ciently in the photocatalytic process. Often a collector is necessary to reach a suitable

energy concentration in the reaction container. The problem of efficient solar collec-

tors is an intensely discussed topic [118, 119]. The parabolic-trough collector (PTC)

[120], and the compound parabolic collector (CPC) [121] are often used photocatalytic

applications. What becomes quickly apparent when dealing with solar collectors is the

necessary space to assemble such structures. Solar driven systems can only be incor-

porated in locations that can provide this space.

Another obvious drawback is the dependence on weather and daylight. Processes

that need to be continuous, for example, urban wastewater treatment or industrial

around the clock production, cannot be covered. This limits the use of solar light for

41

Page 42: Modeling and simulation of photocatalytic degradation of ...

2. Introductionwater cleaning to situations, where the contaminated water can be stored temporarily

overnight and in case of unfavorable weather.

To assess if solar is a suitable solution, the location of a system’s deployment plays

a crucial role. Most parts of central Europe receive just sparse solar irradiation on

average over the year. In particular, during the winter, system performance would

be reduced severely. A comparison between central Europe and south-east Asia in

Fig. 2.10 demonstrates the quantitative difference in available solar energy. While a

solar-based reactor to reduce the concentration of organic dyes in the effluent of the

textile industry in the south of Vietnamwould be sensible, a similar project in Germany

would not be feasible. The best conditions exist in the desert regions around the tropic

circles. Nevertheless, the needs in these parts of the world are generally different from

the possibilities provided by this technology.

2.4.2. Fluorescent tubes and mercury-vapor lamps

As discussed previously, the most common photocatalytic materials need light in the

UV range. Artificial light sources with a wavelength below 400nm are needed to enable

a continuous process. The most famous UV-lamp is the so-called “blacklight.” These

lights are constructed like typical fluorescent tubes. A low-pressuremercury-vapor gas-

discharge lamp forms the base. Under low pressure, the two dominant spectral lines

emitted are at 185 and 254nm. Both lines would be harmful in normal circumstances,

due to the formation of ozone and possible damage to skin and eyes. In the “blacklight”

the emission at 254nm is converted by the phosphors material covering the inner

lamp tube. A typical phosphor that is contained in the lamps used for the experiments

described in this work is strontium tetraborate doped with Europium (SrB4O7:Eu). The

excited phosphor emits a 20nmwidth peak at 372nm [122]. Wavelengths outside the

desired window are filtered so that the tube glows in a deep purple. The energy from

the 185nm peak cannot be used with this kind of lamp and is lost to heat. This lamp

type is useful for experiments because no extensive security measures are needed,

and the wavelength is close to the absorption edge of the conventional catalysts ZnO

and TiO2.

If the phosphor layer is omitted and fused quartz glass used the UV emission of the

low-pressure mercury-vapor lamp can be used directly. As stated above, this kind of

lamp should only be used in a controlled environment. High efficiency of around 35%

42

Page 43: Modeling and simulation of photocatalytic degradation of ...

2.4. Light sourcescan be reached due to the missing conversion step and filters. Most of the energy is

converted into light at a wavelength of 254nm. The ozone generated from the deep

UV-light can be used to create a synergistic effect with the photocatalysis [123].

Overall, mercury vapor based lamps are cheap and effective. Reactor systems based

on these tubes are limited in their design choices because the smallest available com-

mercial diameter is around 15mm.

2.4.3. Light-emitting diodes

Light-emitting diodes (LED) are semiconductor devices that allow electric current to

pass through in only one direction. In a diode, two semiconductors are forming a

p-n-junction. When a LED is operated with a forward bias, recombination occurs at

the interface between the electron-rich (n-type) and electron depleted (p-type) layer

generating light. In conventional diodes, the recombination generates just heat.

H. Round reported the first indication that the production of light from a semiconduc-

tor is possible in 1907 [124]. While studying the diode nature of siliciumcarbid Round

observed a faint yellowish light emerging from his sample. It took nearly 50more years

until, in 1955, R. Braunstein demonstrate infrared light generation with a LED [125].

He linked the observed wavelengths produced by gallium antimony (GaSb), gallium ar-

senide (GaAs), indium phosphide (InP), and germanium-silicon (Ge-Si) alloys directly to

the recombination of the charge carrier and the bandgap of the materials. This lead

finally to the first commercial product, whose ancestors can still be found in remote

controls to this day. The jump to visible light succeeded in 1961 with Ga(As1–xPx) pro-

ducing red light at 710nm [126]. During the 1980s and 90s, the race towards a blue

LED took place. In 2014 Isamu Akasaki, Hiroshi Amano, and Shuji Nakamura received

the Nobel Prize in Physics for their work on for “the invention of efficient blue light-

emitting diodes which has enabled bright and energy-saving white light sources.” The

problematic gallium nitride system was, in the end, the key to a bright blue LED [127].

The development of LEDs in the near UV (300-400nm) and deep UV (200-300nm)

started around 2000 [128]. While the development was mainly driven by the idea to

create an improved white LED through the combination of a near-UV LED with a phos-

phor mixture, today, UV LEDs start to replace the traditional mercury vapor based

solutions even in industrial applications. This trend is accelerated by the “Minamata

Convention on Mercury” from 2013, where many nations pledged a reduced usage

43

Page 44: Modeling and simulation of photocatalytic degradation of ...

2. Introductionand trade of mercury [129]. The success of UV-LED is based on their rapid develop-

ment over the last years. In 2002 the best UV-LED at 365nm achieved an output of

about 0.108W. Ten years later, the output of high powered UV-LEDs was already over

100fold at 12W [128]. The achievable wall-plug efficiency of over 22% at 365nm up

to 77% at 415nm makes LED very attractive for photocatalytic applications [130]. Be-

sides the higher efficiency, LED-solutions exhibit a longer lifetime (in the near UV), a

stable intensity over the product life and are easier to integrate, due to smaller foot-

print and more manageable heat and temperature requirements.

2.4.4. Organic light-emitting diodes

The basic work principle for organic light-emitting diodes (OLED) is close to the LEDs.

Instead of an inorganic semiconductor is the emissive electroluminescent layer formed

by small organicmolecules or polymers. These building blocks enable the use of simple

and flexible substrates and a wide variety of sizes and designs.

OLED could have a significant impact on the design of photocatalytic reactors. The

possibility to stack light generation and photocatalytic material over a wide area with

a minimal thickness could enable new designs. Thin layers of catalyst would be backlit

from a central OLED core.

While the development of near-UV range OLED is on its way, the technology is still not

ready for integration [131, 132, 133, 134]. Looking back on the development history

in the field of LEDs and their struggle with the high energetic blue and UV light, it is

not far-fetched to anticipate the disruptive shift powerful UV-OLEDs could have. At the

moment, the main drawbacks is the energy output and the far too short lifetime of the

modules.

44

Page 45: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methods

3.1. Analytic methods

3.1.1. Nanoparticle characterization

The nanoparticles were characterized for the following parameters: surface area, tex-

ture and diffuse reflectance spectra. The specific surface area of the particles was de-

termined by the Brunauer-Emmett-Teller (BET)method [135]. Thismethod is based on

the theory on multilayer gas adsorption presented by the three name-giving authors

in 1938 [136]. Measurements of the adsorbed gas amount on the external and acces-

sible inner surface under different parameter sets are conducted. With multiple such

measurements, the adsorption isotherm (Eq. 3.1) can be used to extract the specific

amount of gas adsorbed in a monolayer covering the entire surface. The investigated

material must not absorb the chosen measuring gas. That inaccessible pores cannot

be measured with this method is not problematic in this use-case, as pollutants would

not be able to reach these spaces.

pna (p0 – p) = 1

nmCBET +CBET – 1nmCBET · p

p0 (3.1)

The textural properties of the photocatalytic nanoparticles were analyzed by nitrogen

adsorption-desorption at its boiling point (77 K) in a Micromeritics TriStar analyzer (Mi-

cromeritics, Norcross GA). Before performing adsorption experiments, samples (0.5 g)

were outgassed at 26.7 Pa and 350 °C for 6 h.

45

Page 46: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methodsIn the same year as the theoretical foundations of the BETmethod was laid, the first ar-

ticle about the scanning electronmicroscope (SEM) was published [137] with an image

of a ZnO crystal surface. In the SEM, a super fine electron beam is scanning in a raster

over the probe surface. Through the recording of the backscattering and secondary

electrons emanating from the surface, an image can be created. SEM reaches magni-

fication in the order of 100000 which is significantly higher than transmission light mi-

croscopy but lower than high-resolution transmission electron microscope. The mor-

phology for both used particle types was analyzed with an SEM operated at 10 kV and

25 kV.

As the final analytic method for the photocatalytic nanoparticles, diffuse reflectance

spectroscopy was used [138]. This method is based on the theory from Kubelka and

Munk from 1931 aimed towards colored paints [139]. It describes the radiation flux Iinto a body that can both scatter and absorb. Split into infinitesimal small layers of

thickness dx the fluxes in the direction of the initial irradiation I and the opposite di-

rection J is described in Eqs. 3.2 and 3.3.

–dIdx = –(K + S) I + S J (3.2)

dJdx = –(K + S) J + S I (3.3)

A solution for these differential equations was found by Kubelka some years later [140].

The general solution is presented in Eq. 3.4.

R = 1 – Rg [a – b coth(bSt)]a – Rg + b coth(bSt) (3.4)

a = 1 + K/Sb = √a2 – 1

In the solution, t is the thickness of the sample, and Rg is the reflectance of the back-ground material the sample is placed upon. In practice, the samples are beyond a

critical thickness which renders the background obsolete. For t → ∞ the hyperbolic

cotangent approaches 1 and Rg = 0. The resulting Eq. 3.5 is also known as Kubelka–Munk function FK,M.

46

Page 47: Modeling and simulation of photocatalytic degradation of ...

3.1. Analytic methods

FK,M(R∞) = KS = (1 – R∞)2

2R∞(3.5)

In the case that the scattering is not depending on the wavelength, FK,M(R∞) is propor-

tional to the actual absorption spectrum. For samples with grain sizes over 1.5 µm, it

was shown that this assumption is true [141]. In the experiments with smaller grain

sizes the scattering coefficient increases significantly in the UV region. That results in

a damped response the further the measurement reaches into the UV and the peak

of a measurement is shifted to shorter wavelengths.

The spectra for the material used were obtained with an UV-vis spectrophotometer

2101PC (Shimadzu) in the range of 190 to 600nm. It was equipped with a diffuse

reflectance attachment and a reference standards disks made from pressed barium

sulfate (BaSO4).

3.1.2. Ultraviolet-visible absorption spectrometry

Absorption spectroscopy is widely used to quantify the concentration of a large num-

ber of different molecules. The spectra are typically measured for wavelengths in the

range from 200 to 800nm. Depending on the molecule, a subset can be selected to

speed up the measuring process.

The method is based on the Beer–Lambert’s law (Eq. 3.6) describing the connection

between transmittance T , sample thickness t, and sample concentration Cs.

A = – log TL = logIiI = εtCs (3.6)

This equation can be used directly if the sample could be held in a perfect transparent

container. In practice, the light beam has to pass through the air as well as the cuvette

material and cross the interfaces between air, cuvette, and sample twice. To account

for these effects, the intensity of the sample Isample is measured in relation to the in-

tensity of the beam passing through a comparable setup filled with the solvent only

Isolvent. With Eq. 3.7, a close approximation of the actual absorbance and transmittance

47

Page 48: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methodsis calculated [142].

IsolventIsample ≈ IiI (3.7)

For most organic molecules, the relation between the concentration and the observed

absorbance is linear, following Beer–Lambert’s law. However, this linear behavior is

limited to low concentrations (< 0.01mol L−1). At high concentrations, themolecules do

not act isolated from each other anymore, and the absorption of light in one molecule

can influence the state of its neighbors. Some molecules exhibit non-linearity already

at lower concentrations. The cation of methylene blue, that is used as an example

molecule for dyes in this work, does not show adherence to Beer–Lambert’s law even

at a concentration under 1 • 10−6mol L−1. In this case, a densely populated calibration

curve is used, to not rely on linear dependency [143].

In all experiments, the samples were analyzed using a CARY-100 UV-vis (Varian) spec-

trophotometer. The concentrations were determine by comparing the measured ab-

sorption spectra to the calibration curves.

3.1.3. SPE-HPLC-MS/MS

UV-vis spectrometry is limited to simple systems with a low variety of substances. The

effluent of a sewage treatment plant contains a high number of differentmolecules in a

single sample. Besides, the low concentrations of these substances provides a further

challenge. Therefore, R. Gurke et al. developed a novel method at the Institute of Clin-

ical Pharmacology in Dresden, which is based on three main techniques [144]. In the

first step, a sample is prepared using solid-phase extraction (SPE). By letting the sam-

ple run through a solid phase that the solvent can pass unimpeded, the investigated

substances are adsorbed, and the molecules of interest are enriched. This allows for

the analysis of very low concentrated samples. In the next phase high-performance

liquid chromatography (HPLC) is applied. Thereby, the molecules are spread based on

the interaction with the material in the columns. Through this process, the following

analytic method is utilized more efficiently. The concentration is then measured by a

tandemmass spectrometry (MS/MS). The combination of two mass spectra allows the

first to preselect a mass range from the fraction generated by the HPLC for detailed

48

Page 49: Modeling and simulation of photocatalytic degradation of ...

3.1. Analytic methodsanalysis in the second one.

The developed SPE-HPLC-MS/MS-method is described in detail in Gurke et al. 2015

[144]. The following summary is based on the joint publication with R. Gurke [145].

From the solutions generated in the experiments 1mL of the sample was taken, ad-

justed to a pH of 3 by using formic acid and spiked with 100 µL of the internal stan-

dard (IS) solution (10µg L−1). The samples were extracted using an Abimed ASPEC XL

(Gilson) withOasis HLB 10mg Extraction Cartridges (Waters). Under a gentle air stream

at 50 °C the eluates were evaporated to dryness and re-dissolved in 250 µL mixture of

solvent A and solvent B (80/20, v/v). The solvents A (97/3/0.05; v/v/v) and B (5/95/0.05;

v/v/v) were a composition of 2mmol L−1 ammonium acetate solution, acetonitrile, and

formic acid. An LC-MS/MS system, consisting of a Dionex-HPLC composed of an Ul-

tiMate3000 Pump and Autosampler (Thermo Fischer Scientific) with a Chromeleon 7

Chromatography Data System (Dionex Softron) and coupled to an API 4000 tandem

mass spectrometer (AB Sciex) equipped with an electrospray ionization source (ESI),

was used for the analyses of the samples. The chromatographic separation was per-

formed with a Synergi 2.5u HydroRP 100A, 100 x 2.0mm column and a C18 security

guard 4mm x 2mm (both Phenomenex) using a multi-step gradient out of solvent A

and B with a total runtime of 15min. For the analyses, an injection volume of 20 µL

was chosen. The mass spectrometric analyses were performed in multiple reaction

monitoring (MRM)mode with positive electrospray ionization. The Analyst data system

version 1.6 (AB Sciex) was applied to control the MS, evaluate the peak area, analyze

the regression of calibration curves, and calculate the final concentrations.

3.1.4. Non-purgeable organic carbon

To analyze the performance of a photocatalytic system, the concentration evolution

of the initial molecule is not sufficient enough. The tracking of all intermediates would

be the preferred solution but is practically impossible for complex organic molecules.

However, measuring the overall amount of organic carbon instead allows for monitor-

ing the degradation process in its entirety.

To limit the measured carbon to the organic fraction, solved inorganic carbon need to

be removed. As shown in Reac. 7, the carbonic acid reaction can be moved towards

dissolved CO2 by acidifying the solution. Through bubbling the solution with O2, the

inorganic carbon is purged.

49

Page 50: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methods

CO2 (aq) + H2O(l) H+(aq) + HCO

–3 (aq) {7}

Subsequently, the remaining non-purgeable organic carbon (NPOC) is quantified [142]

using a multi N/C 3100 from jena analytic. The sample is broken down, by injecting

it into a combustion tube, that is filled with catalyst and heated to 800 °C. A non-

dispersive infrared sensor measures the produced CO2 concentration. The carbon

dioxide concentration integrated over time can be related to the carbon content in

the sample.

3.2. Experimental investigations

3.2.1. Model substances

The antibiotic ciprofloxacin, and the dye methylene blue were selected as model sub-

stances. A core requirement for the selection was a significant peak in its absorption

spectra for the use with an UV-vis absorption spectrometer.

Ciprofloxacin (C17H18FN3O3) is a synthetic antibiotic from the fluoroquinolone group

developed by Bayer. This antibiotic is most effective against Gram-negative bacteria.

Its bactericidal effect is based on the inhibition of enzymes essential for cell division.

While themedical indications to use this drug are reduced significantly in the last years

through to newly discovered side effect, it is still in use for a wide area of bacterial

infections. A typical dose is 1 g d−1, that results in a serum concentration of around

2.5mgL−1. With a typical half-life of 4 - 7 h, the drug is excreted over the kidneys.

Over 69% of the given dose leaves the body unaltered (44.7% urine, 25.0% feaces).

Another 19% are discharged as close metabolites [146]. When the drug reaches the

wastewater system, it is diluted significantly but is still quantifiable. In the influent of

the Dresden wastewater treatment plant (WWTP) a concentration of 422ng L−1 was

measured [147]. As stated before, the removal of antibiotics from our wastewater is

essential to prevent the development of resistances against them. Mutations in the

target enzymes, topoisomerase II and IV can result in resistance against ciprofloxacin.

Multiple simultaneous mutations can lead to a more broad clinical resistance against

the whole class of similar antibiotics. The presence of antibiotics in the environment

50

Page 51: Modeling and simulation of photocatalytic degradation of ...

3.2. Experimental investigationscan apply a selection pressure on bacteria even outside of a therapeutic context.

Methylene blue (C16H18ClN3S) was selected as an example of an organic dye. This syn-

thetic cationic dye was first produced in 1876. The deep blue color is mostly used to

dye cotton in the textile industry. It is still in use today, even though the high number of

available dye variants naturally limits its use. While the textile industry consumes the

most substantial portion of methylene blue, its interaction with tissue is the main rea-

son for its detailed scientific examination. Very early methylene blue was already used

to stain biological systems. The discovery of its strong interactions with pathogens led

to its use as an anti-malarial drug in 1891. Methylene blue can, therefore, be seen

as the first synthetic drug. Its selective affinity for nervous tissue made it useful in

the treatment of neuropsychiatric disorders. A great variety of drugs were developed

based on methylene blue [148, 149]. Due to its prominence in the biological field, it

was a natural choice early on as a model substance. Over the decades, it became a

standard for testing photocatalytic degradation processes.

3.2.2. Adsorption-desorption

To understand the dynamics between the adsorption and desorption, experiments

conducted in darkness are necessary. The entire reaction vessel was sleeved with

aluminum foil, and the laboratory was just scarcely illuminated to limit the possible

introduction of light during the sampling procedure.

In the initial experiments 60mL of prepared solution of ciprofloxacin (12.5 µmol L−1)

and methylene blue (11.0 µmol L−1) were filled in 100mL beakers (VWR) and continu-

ous stirred. To those solutions, 1 g L−1 of TiO2 P25 (Evonik) or ZnO (IOLITEC Ionic Liq-

uids Technologies GmbH) was added. The first sample was taken with a syringe 40 sec-

onds after the photocatalytic nanoparticles were mixed in with the solutions. Multiple

further samples were taken over the next half hour every few minutes. The nanoparti-

cles need to be removed to stop further adsorption. The sampleswere passed through

a fine filter (Rotilabo nylon, pore size 0.2 µm). In preparation for the following UV-vis

analytic, the remaining nanoparticles were removed by double centrifugation for 60

minutes. The final supernatant was then measured.

As discussed further in the introduction to the single organic species model (p. 75),

the first measurement is the most important to determine adsorption constant (Eq.

3.18). The experiment was repeated additional five to ten times for each combination

51

Page 52: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methodsof material and catalyst to improve the reliability of the measurements. In these addi-

tional experiments just the initial concentration and the result after 40 seconds was

recorded.

3.2.3. Photocatalytic degradation

The degradation experiments were carried out in 100mL beakers from VWR with a

5 cm diameter. They are made from borosilicate glass with a wall-thickness of 3.3mm

and are therefore well suited for experiments involving UV-light. The illuminating de-

vice, manufactured by UMEX, was equipped with 6 Philips 8W mercury fluorescent

tubes. Two beakers were placed in 15 cm distance from the illumination device, with

an equal distance away from the center. The UV intensity inside the beakers ranged

from 18 to 19Wm−2 and was determined by a UV34 Lux Meter (PCE). The full setup is

shown in Fig. 3.1.

Figure 3.1.: Experimental setup for the photocatalytic degradation experiments - the

beaker on the right side does not contain photocatalyst and serves as a

control.

Prior to the start of the degradation experiments, samples were stirred in the dark for

at least 30 minutes. Thereby, the solution reaches the adsorption-desorption equilib-

rium between the organic molecules on the photocatalyst surface and in the solution.

An amount of 50mL of this solution was then exposed under continuous stirring to UV-

radiation with an intensity peak at 365nm. The specific spectrum of the light source is

incorporated in Fig. 4.1.

52

Page 53: Modeling and simulation of photocatalytic degradation of ...

3.2. Experimental investigationsSamples were taken from the solution at multiple points in time to analyzing the tem-

poral degradation of organicmolecules. The same filtering and centrifugation steps, as

described in the adsorption-desorption experiments, were perfomed to remove the

catalyst nanoparticles. Concentration measurements were again conducted by UV-vis

absorption spectrometry.

3.2.4. Wastewater treatment plant effluent

The influence of a chemical background and ultra-low concentration was examined to

test the model under more realistic circumstances. For this experiment, the effluent

of a wastewater treatment plant was used. The Stadtentwässerung Dresden GmbH

provided samples from the municipal plant in Kaditz (Fig. 3.2). In this plant 55 • 106m3

sewage water is treated per year. This treatment plant currently cleans the sewage

of 650000 people and has a design capacity of 740000 inhabitant equivalents. The

structure of the WWTP consists of a primary clarifier, an activated sludge reactor, and

a secondary clarifier [150]. Water was collected as a 24h flow proportional composite

effluent sample on June 24th, 2014, after two relative dry weeks (4.2 Lm−2) and stored

at 4 °C. Standard measurements of the effluent sample are summarized in Tab. 3.1.

All effluent samples were filtered by a filter paper (VWR pore sizes 5-13 µm) to remove

suspended particulate matter.

Figure 3.2.: Wastewater treatment plant in Kaditz operated by Stadtentwässerung

Dresden GmbH.

Besides the collected effluent sample, amixture ofMillipore water with carbamazepine

(12mgL−1) was created. From both samples 50mL were mixed with ZnO and TiO2

(1 g L−1) and exposed under continuous stirring to UV-radiation. The samples were

53

Page 54: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methodsfiltered and centrifuged as described previously. For the spiked sample, the measure-

ments were performed with the Varian CARY-100 UV-vis spectrophotometer. The un-

altered effluent sample was analyzed using SPE-HPLC-MS/MS.

Table 3.1.: Standard parametersmeasured in the effluent of thewastewater treatment

plant in Kaditz. Taken as 24h flow proportional composite effluent sample

on June 24th, 2014

chemical oxygen demand (COD) 37 mgL−1

biochemical oxygen demand (BOD) 4 mgL−1

Nitrogen

Ntotal 12.0 mgL−1

total Kjeldahl nitrogen (TKN) < 5.0 mgL−1

Nammonium 0.31mgL−1

Nnitrite 0.03mgL−1

Nnitrate 7.40mgL−1

Ninorganic 7.74mgL−1

Phosphor

Ptotal 0.86mgL−1

Nphosphate 0.56mgL−1

pH 7.5

The initial experiment was limited to one pharmaceutical molecule. Eventually, mul-

tiple drugs should be monitored at the same time in the next step. Target pharma-

ceuticals were preselected based on their prescription statistics. This selection of 55

molecules was based on a study by Gurke et al. [144]. For monitoring the degradation

process, the initial concentration (Ci) of the pharmaceuticals needs to be significantlyhigher than the corresponding detection limit. Therefore, we set the lower limit to be

Ci > 0.3µg L−1.After an initial analysis of the effluent sample, 14 pharmaceuticals were selected to

be monitored in the degradation experiment. Based on the initial concentration crite-

ria the anticonvulsants carbamazepine Ci = 1.29µg L−1, gabapentin Ci = 11.30µg L−1,lamotrigine Ci = 0.98µg L−1, and oxcarbazepine Ci = 0.63µg L−1, the antidepressant

venlafaxine Ci = 0.58µg L−1, the beta blockers bisoprolol Ci = 0.58µg L−1, celiprolol

Ci = 0.35µg L−1, and talinolol Ci = 0.43µg L−1, the lipid-lowering drug bezafibrate Ci =0.48µg L−1, the opioid analgesic tramadol Ci = 0.624µg L−1, as well as the angiotensinreceptor antagonists candesartan Ci = 1.30µg L−1, eprosartan Ci = 0.56µg L−1, irbesar-

54

Page 55: Modeling and simulation of photocatalytic degradation of ...

3.2. Experimental investigationstan Ci = 1.50µg L−1, and valsartan Ci = 3.59µg L−1 were analysed in the degradation

experiment. The selected pharmaceuticals are listed with their corresponding lower

limit of quantification (LLoQ) in Tab. 3.2. It is essential to keep in mind that this data is

a mere snapshot because the concentration of micropollutants can significantly vary

in sewage samples based on a diverse range of parameters, for example, weather

conditions.

Table 3.2.: Selected pharmaceuticals from different drug classes to be monitored in

the degradation experiment. in themost left column the lower limit of quan-

tification (LLoQ) is listed.

analyte drug class provider LLoQ in ng L−1

Carbamazepine anticonvulsant Sigma 50

Gabapentin Pfizer 200

Lamotrigine Sigma 50

Oxcarbazepine Cerilliant 50

Venlafaxine antidepressant Wyeth 50

Bisoprolol Merck 50

Celiprolol beta blocker Oxprenolol 50

Talinolol LGC Standards 50

Bezafibrate lipid-lowering drug Sigma 50

Tramadol opioid analgesic Sigma 50

Candesartan AstraZeneca 50

Eprosartan angiotensin receptor Sigma 50

Irbesartan antagonist Sigma 50

Valsartan Sigma 100

As the SPE-HPLC-MS/MSmethodneeded significantlymore volumeper sample tomea-

sure all the concentrations, the sample volume was increased from 50mL to 100mL.

The treated wastewater were mixed with photocatalyst until a concentration of 1 g L−1

was reached. The altered setup lowered the UV intensity inside the bigger beakers to

15-16Wm−2 as determined by the UV34 LuxMeter (PCE). After the suspension was ex-

posed under continuous stirring to the UV-radiation aliquots of 2mL were withdrawn

at specific time intervals. Because the sample was analyzed by the SPE-HPLC-MS/MS

method, the filtering was skipped, and the samples were centrifuged for 60minutes to

remove the significant part of catalyst particles. As controls, the experiments were in

addition carried out in the absence of catalyst nanoparticles or without UV irradiation.

55

Page 56: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methodsAll other parameters in the control experiments were kept unchanged.

3.3. Modeling approach

This sectionwill focus on the theoreticalmodel of the degradation of organicmolecules

through photocatalysis. After describing the fundamental interactions in the system,

the model variants are described in detail. Beginning with the single-species model

and then continuing with the three alterations of the multi-species model. The pro-

cesses occurring during the photocatalytic degradation of organic components in the

solution are illustrated in Fig. 3.3. The steps towards complete mineralization of the

organic molecules is considered to take place mainly at the surface of the photocat-

alytic material. As radicals are extremely short-lived, the possible reactions of radicals

and organic compounds in the water surrounding the particle are neglected.

The organic components diffuse to the surface of the photocatalytic nanoparticles,

where they are adsorbed andmineralized. Organicmolecules desorb and diffuse away

from the particle at the same time, limiting the mineralization process. This process

is repeated for the fragments of the initial molecule until a complete mineralization

ist reached. For this modeling approach, it is presumed that the oxygen supply for

mineralization is sufficient during the entire process. This can be accomplished by

intense mixing in contact with the open atmosphere or with the help of a bubbler.

For the initial Gedankenexperiment, we are starting from a homogeneous solution of

the organic molecules. As soon as the photocatalytic material is added to the system,

the organic molecules start to be adsorbed at the nanoparticle surface. This adsorp-

tion results in a slight concentration gradient around the particles, depending on the

adsorption rate. However, the concentration profile strongly flattens in a brief mo-

ment due to the relatively fast diffusion of the organic molecules in the solution. In the

dark, without a reaction flow, the desorption and diffusion will finally lead again to a

homogeneous solution with a reduced concentration. When the reaction is initiated

through light, the mass-flow towards the surface will continue. To assess if diffusion

processes in the solution need to be taken into account, an estimation of the diffusion

speed of the model substance ciprofloxacin is done.

56

Page 57: Modeling and simulation of photocatalytic degradation of ...

3.3. Modeling approach

jdes

jads

jreac

M

photocatalytic material initial molecules fragments

Figure 3.3.: Illustration of the processes occurring during the mineralization of organic

molecules at the surface of photocatalytic nanoparticles in solution. Due to

degradation and desorption, various intermediate organic molecules (frag-

ments) can emerge on the surface and in the solution. The degradation

steps on the surface create the final mineralized products (M).

D = kBT/(6π η r) (3.8)

By means of the Stokes-Einstein equation (Eq. 3.8) where η = 10−3 Pa s is the dynamicviscosity of water at 20 °C and r = 0.46nm is the radius of a sphere with a volume

equal to the van-der-Waals volume of ciprofloxacin (0.41 nm3), we find a diffusion co-

efficient of 4.65 • 10−10m2 s−1. A similar value of around 4.65 • 10−10m2 s−1 is obtained

when using the empirical equation presented by Wilke and Chang [151]. The order

of magnitude of the maximal diffusion distance between the nanoparticles, which the

molecules have to overcome, is given by a = C–1/3P

, where CP is the particle concen-tration. For a particle suspension of 1 g L−1 TiO2 and an upper estimate of the particle

radius of 50nm (considering possible agglomeration), one finds a ≈ 1300nm. This

results in a characteristic diffusion time tD = a2/D of 3.6ms [152]. Since this time is

57

Page 58: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methodsorders of magnitude smaller than the time, where significant mineralization of the or-

ganic molecules occurs, a homogeneous concentration of organic molecules between

the nanoparticles can be assumed in the following model approaches.

3.3.1. Single organic species model

The first modeling approach follows previous studies (see, for example, [45, 35] and

references therein) and consider an idealized model system with only one organic

species in the aqueous solution. Organic molecules in the solution adsorb onto the

particle and are mineralized due to photocatalytic reactions. The molecules can also

desorb before the mineralization. These three possible paths are shown in the reac-

tion equation below (Reac. 8).

A(aq) A(ad) M(aq) {8}

In the first part of the experiment that is performed in the dark, only the adsorption

and desorption plays a role. Temporal changes of the molecule concentration in the

solution CA(aq), in m−3, and the concentration on the photocatalyst particle surface

CA(ad), in m−2, are given by Eqs. 3.9 and 3.10.

d

dtCA(ad) = jads – jdes – jreac (3.9)

d

dtCA(aq) = as(jdes – jads) (3.10)

These equations include the specific surface area of the nanoparticles aS and the threedifferent molecule fluxes as shown in Eqs. 3.11-3.13.

jads = kads (1 – Θ)CA(aq) (3.11)

jdes = kdes CA(ad) (3.12)

jreac = kreac CA(ad) (3.13)

58

Page 59: Modeling and simulation of photocatalytic degradation of ...

3.3. Modeling approachThe adsorption flux jads is proportional to the molecule concentration in the solutionand the free surface area of the particles. It is assumed that a monolayer is formed.

The surface coverage Θ is given by Eq. 3.14.

Θ = Am CA(ad) (3.14)

In this equation, Am is the surface area covered by one molecule. The largest possible

projection of the solvent-accessible surface is used to estimate this area [153]. For

this parameter, the van der Waals surface is probed by a molecule. To reduce the

resource demanded to generate such a surface, the solvent is abstracted as a sphere.

Most commonly, water is used as the probe molecule with a corresponding sphere

radius of r = 1.4 Å.

a) b) c)

Figure 3.4.: Steps to calculate the maximal projected accessible surface area of cipro-

floxacin: a) spatial structure generation, b) rotation and casting circles with

combined van der Waals and probe radius, and c) eroding with probe disk.

For this model, the full three-dimensional surface of the molecule is not needed, as

we are only interested in the maximal covered area. To efficiently calculate Am the

following steps are performed as depicted in Fig. 3.4. First, the spatial structure of the

organic molecule is gathered from a central repository like PubChem. For molecules

that are not yet included in a database, the 3D optimization can be done with the

MMFF94 force field [154]. Merck developed this force field with organic and drug-like

59

Page 60: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methodsmolecules in mind. Therefore, the list of supported elements is limited to the ones

typically needed for organic molecules (C, H, N, O, F, Si, P, S, Cl, Br, I). The 3D structure

of ciprofloxacin is shown as an example in Fig. 3.4 a.

In the second step, the principal axes of the molecule are determined based on the

atom positions. Knowing the principal axis enables the rotation of the molecule in

such a way that the maximal area is stretched out along the x and y-axis. Now the

problem is transformed into a two-dimensional system by omitting the z-axis informa-

tion. Centered on all atoms in the molecule circles are drawn with the corresponding

van da Waals radius combined with the probe radius (Fig. 3.4 b). The van der Waals

radii are taken from the 1964 publication by Bondi [155]. An exception is hydrogen,

which radius was found to be overestimated by 0.1 Å by Rowland and Taylor in 1996

[156]. All radii that are not listed in these publications are assigned the value of 2.0 Å.

However, these elements do not play a crucial role in organic molecules.

In the last step, the newly created area is eroded using a disk with the same radius

as the probe. The resulting area (Fig. 3.4 c) is used to determine Am. It is important tonote that the resulting value estimates the maximal surface area and can not express

the real adsorption geometries of organic molecules.

The reaction flux jreac describes the number of mineralized molecules per area and

time. Both desorption and reaction fluxes are proportional to the concentration of

organic molecules on the surface. In the limiting case of small surface coverage Θ � 1,

the system of ordinary differential equations, Eqs. 3.9 & 3.10, becomes linear and can

be solved analytically. The full derivation is presented in Appendix - chapter A, and the

solution is shown in Eqs. 3.15 and 3.16.

CA(aq)(t) = k1 exp(λ1 t) + k2 exp(λ2 t) (3.15)

CA(ad)(t) = k1λ1 + as kadsas kdes exp(λ1 t) + k2 λ2 + as kadsas kdes exp(λ2 t) (3.16)

k1 = as kdes CA(ad),0 – (as kads + λ2)CA(aq),0λ1 – λ2k2 = as kdes CA(ad),0 – (as kads + λ1)CA(aq),0λ2 – λ1

60

Page 61: Modeling and simulation of photocatalytic degradation of ...

3.3. Modeling approach

λ1 = –1

2

(θ0 +

√θ20 – 4as kadskreac)

λ2 = –1

2

(θ0 –

√θ20 – 4as kadskreac)

θ0 = as kads + kdes + kreac

The rate constants kads, kdes, and kreac are derived experimentally. First, the adsorptionand desorption rates are obtained from experiments of the concentration evolution

CA(aq)(t) in the dark. After an initial decrease of the concentration, a stationary value

is reached due to an adsorption-desorption equilibrium at the particle surface. For

sufficiently diluted concentrations, the emerging coverage on the particle surface is

minimal: Θ � 1. The concentration evolution results as solution of Eq. 3.9 with jreac =0 as

CA(aq)(t) = CA(aq),0exp[–(kads as + kdes)t] + kdes(kads as)–11 + kdes(kads as)–1 (3.17)

Generally, the adsorption rate constants can be determined from accurate measure-

ments of the initial concentration drop dCA(aq)/dt at t = 0, when the coverage is still

negligible

kads = –CA(aq)(0)as CA(aq),0 (3.18)

With the knowledge of kads, the desorption rate constant kdes is obtained from themea-

sured asymptotic concentration value CA(aq),∞ = CA(aq)(t → ∞). From the adsorption-

desorption equilibrium jads = jdes, we find with Θ = AmCA(ad)

kdes = kads(1 – AmCA(ad),∞)CA(aq),∞CA(ad),∞ (3.19)

61

Page 62: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methodswhere the asymptotic surface concentration follows from the total particle balance as

CA(ad),∞ = 1

as (CA(aq),0 – CA(aq),∞). (3.20)

After determination of the rate constants kads and kdes by means of experiments con-ducted in the dark, the reaction rate constant kreac is determined from concentration

measurements obtained in the degradation experiments under illumination. During

the fitting procedure, Eqs. 3.9 and 3.10 are solved numerically.

For the case of a fast establishing adsorption-desorption equilibrium with a quasi-

stationary surface concentration CA(ad) = 0, the surface concentration in the system

of equations 3.9 and 3.10 can be eliminated. This would be the case when |CA(ad)| �kads CA(aq) or kdes CA(ad). The solution of Eq. 3.9 in the limit of small coverage Θ � 1

reads then CA(ad) = CA(aq) kads(kdes + kreac)–1. Insertion of this expression into Eq. 3.10yields CA(aq) = –kappCA(aq) with the solution

CA(aq)(t) = CA(aq),0 exp(–kappt) (3.21)

where the apparent degradation rate constant is given by

kapp = askadskreackdes + kreac . (3.22)

Thus, for kdes � kreac, we find kapp = askads. The degradation rate becomes adsorption-limited and does not depend on the reaction rate constant. In the opposite case,

kdes � kreac, one obtains kapp = askadskreac/kdes. The degradation rate is proportionalto the reaction rate constant.

62

Page 63: Modeling and simulation of photocatalytic degradation of ...

3.3. Modeling approach

3.3.2. Multiple organic species model

The single organic species model is an appropriate tool for analyzing the adsorption-

desorption process of the initial organic compounds. However, photocatalytic miner-

alization of organic molecules is a far more complicated process, accompanied by the

formation of various intermediates in several reaction steps. Therefore, two idealized

mechanisms were considered to describe the mineralization reaction in more detail

(Fig. 3.5). In the first model, the initial organic molecule is oxidized step by step. It

can be imagined that in this model, the mineralization occurs just at the outer ends

of the initial molecule and its intermediates. Every time a reaction step happens, only

a small part (e.g. one carbon atom) is oxidized. This model is thus referred to as the

incremental oxidation model.In the second model, referred to as fragmentation model, it is considered that bondswithin the molecule are destroyed due to photocatalytic reactions. In contrast to the

first model, molecules can now break into two smaller fragments without mineraliza-

tion.

Based on the secondmechanism, themodel is extended further. As the initialmolecule

and the intermediates are portrayed as a chain, information from the original molecule

is lost. Through the introduction of excess bonds, the backbone of the base moleculecan be better captured. A reaction step can reduce these excess bonds in the same

way the fragments are broken apart.

Intermediates may desorb from the catalyst surface instead of further undergoing di-

rect reactions in all models. Thus concentrations of different species need to be intro-

duced in the solution and on the surface of the catalyst. The different species will be

characterized solely by their size.

Incremental oxidation model

The initial organic compounds and their intermediates are described by the number of

carbon atoms n in their structure to build a model that incorporates different species.Within the incremental oxidation model, the molecule degradation is described in an

idealized manner by oxidation and removal of one carbon atom at every reaction step

(see Fig. 3.5). With the separation of each carbon atom, a corresponding fraction of

the remaining molecule is considered mineralized. Adsorption, desorption, and min-

63

Page 64: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methods

C +

+

+

+

C C C C C

C C C C

C C C C C

C

C C

C C

C C C

C C

C C C C

C

C C

C C

C

M

M

M

M M

+

+

+

+

+

C C C C C C C

C C C C C C C

C C C C C C C

C C C C

C C C C

+

+

C C C

C C

C C C C

C

C C

C C

C

C C C

C C C

C C

M

C C

M

C C

. . .

. . .

fragmentation

incremental

Figure 3.5.: Illustration of the reaction steps in the two basic multiple organic species

models: (top) incremental oxidation and removal of single carbon atoms

and (bottom) formation of fragments of molecules by intermolecular bond

breaking and subsequent oxidation of the smallest fragments.

eralization of a molecule of size n can, therefore, be expressed by Reac. 9.

An (aq) An (ad) A(n–1) (ad) + M(aq) {9}

The concentration evolution of molecules of size n ≥ 2 and of the mineralized compo-

nent are described by Eqs. 3.23-3.25.

d

dtCAn(ad) = jn,ads – jn,des – jn,reac + j(n+1),reac (3.23)

d

dtCAn(aq) = as(jn,des – jn,ads) (3.24)

d

dtCM(aq) = as(2j2,reac +

nmax∑n=3 jn,reac

)(3.25)

64

Page 65: Modeling and simulation of photocatalytic degradation of ...

3.3. Modeling approachThe corresponding fluxes are defined in Eqs. 3.26-3.28.

jn,ads = kn,ads (1 – Θ)CAn(aq) (3.26)

jn,des = kn,des CAn(ad) (3.27)

jn,reac = kn,reac CAn(ad) (3.28)

The flux j(n+1),reac in Eq. 3.23 describes the formation of molecules of size n due to

oxidation of molecules of size n + 1. To track the concentration of the mineralized

component CM in the solution, the mineralization fluxes of all molecules is summed as

shown in Eq. 3.25. For simplicity, we assume the rate constants in Eqs. 3.29-3.31 to

vary only with the molecule size n.

kn,ads = κDn (3.29)

kn,des = ν0(n) exp(–Edes(n)kB T

)(3.30)

kn,reac = I0 φ An (3.31)

It is reasonable to assume that the size of the fragments considerably affect the rate

constants. Especially, for chain like molecules with equal repeating units. In general,

the rate constants depend also on the specific element composition and configuration

of the molecules. Roughly, we choose the adsorption rate constant to be proportional

to the diffusion constant of the molecules, which in turn is determined by their size

according to the Stokes-Einstein equation (Eq. 3.8). The molar volume, used to esti-

mate the spherical molecule radius, is supposed to be proportional to the number

of carbon atoms n. Dividing the initial molecule volume up for the intermediates willunderestimate their values. The reaction rate constant is chosen to be proportional

to the area of the photocatalyst surface covered by the molecule. This choice arises

from the idea that the reactions can happen with a constant probability over the en-

tire catalyst surface. If an intermediate covers more of this surface, it is more likely to

encounter a reaction in the same time frame. The remaining constants κ, and φ in Eqs.3.29-3.31 are chosen to match the experimentally found rate constants of the initial

molecule. Little is known about the dependence of the desorption rate constant on

65

Page 66: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methodsthe molecule size for aqueous solutions. For the desorption of organic molecules in

vacuum, temperature programmed desorption studies [157] and molecular dynamics

simulations [158] have been performed. For example, for the desorption of alkanes

CnH2n+2 from the MgO(100) surface, experimental investigations in [157] showed the

desorption energy Edes to depend linearly on the chain length n.

Edes(n) = (6.5 + 7.1n) kJmol–1 (3.32)

The pre-exponential factor ν0 strongly increased in the same setup from 1013.1 to

1019.1 for n varying from 1 to 10. To what extent similar tendencies apply to desorp-

tion in aqueous solutions is not known. We consider, as one assumption in this model,

that similar behavior occurs, where the desorption energy and pre-exponential factor

depend on the molecule size in the following way.

Edes(n) = E0 + E1 n (3.33)

ν0(n) = α0 10α1n (3.34)

The parameters in these equations were chosen so that the desorption rate agrees

with the rate derived fromexperiments in the darkwith the initial organicmolecules. As

a consequence of such a strong size dependence, small intermediate molecules exhibithigh desorption rates, leading to a low mineralization rate of these components. As

demonstrated later in the section 4.3.2 (p. 83), this results in a prolonged decay of the

total organic carbon (TOC) signal even after extended exposure time. Such residual

long-term TOC signals have been reported in [159, 160, 161, 162].

The desorption rate of intermediates depends, of course, not only on themolecule size

but crucially on their chemical structure as well. The TOC signal in many degradation

studies has been observed to vanish entirely after moderate illumination exposure

[163]. For this reason, we consider the case of weak size dependence of the desorptionrate of intermediates.

66

Page 67: Modeling and simulation of photocatalytic degradation of ...

3.3. Modeling approach

kn,des = β0 + β1/n (3.35)

As a possible model function for this size dependence, we have chosen a simple rela-

tionship with only two parameters. The simplicity of the relationship in equation 3.35

facilitates parameter fits from experimental data.

Fragmentation model

In this model, we consider the limiting case that photocatalytic reactions lead to the

progressive destruction of bonds in the molecules on the catalyst surface (Fig. 3.5).

Only the smallest possible fragments are considered mineralized. The initial molecule

is abstracted as a chain, where the number of possible fragmentations of a molecule

of size n is equal to n-1. The corresponding reaction equations for the species An isgiven in Reac. 10.

An (aq) An (ad) A(n–1) (ad) + M(aq) {10}

A(n–2) (ad) + A2(aq)A(n–3) (ad) + A3(aq)A(n–4) (ad) + A4(aq)

· · ·

M(aq) + A(n–1) (ad)

The concentration evolution for n ≥ 2 is described by equations that are similar to the

ones for the incremental model (Eqs. 3.23-3.25).

d

dtCAn(ad) = jn,ads – jn,des – jn,reac + j∗n,reac (3.36)

d

dtCAn(aq) = as(jn,des – jn,ads) (3.37)

d

dtCM(aq) = as j∗1,reac (3.38)

67

Page 68: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methodsHowever, according to Reac. 10, several reaction pathways are possible for an ad-

sorbed molecule. Thus, the evolution of an intermediate molecule of size n is influ-

enced by the reactions of all larger molecules. This is included in the reaction flux

j∗n,reac as the sum over all possible fragmentations of larger molecules.

j∗n,reac = nmax∑m=n+1

2m–1 km,reac CAm(ad) (3.39)

Thereby, assuming m-1 possible fragmentations of a molecule of size m occur with

equal probability, leading to the weight factor 2/(m-1).

Excess bonds model

In the last variant of the multiple organic species model, we include information about

the bonds in the initial molecule. Besides the carbon atoms count, the number of

bonds that exceeds the number of links in the abstracted chain is used. In the case

of ciprofloxacin, there are 17 carbon atoms, and the molecule backbone is build up

by 27 bonds, as shown in Fig. 3.6. While connections to hetero-atoms that are part

of the backbone are counted, bonds to outward-facing hetero-atoms are excluded.

For the example of ciprofloxacin, we would have 16 links in the model chain and 11

excess bonds. While keeping the generalization of the fragmentation model, we now

allow a reaction step that leads just to the destruction of an excess bond. The possible

reactions are summarized below.

AnEb (aq) AnEb (ad) AnE(b–1) (ad) {11}

A(n–1)Eb (ad) + M(aq)

A(n–2)Eβ(n–2) (ad) + A2E(b–β(n–2))(ad)A(n–3)Eβ(n–3) (ad) + A3E(b–β(n–3))(ad)A(n–4)Eβ(n–4) (ad) + A4E(b–β(n–4))(ad)

· · ·

M(aq) + A(n–1)Eb (ad)

68

Page 69: Modeling and simulation of photocatalytic degradation of ...

3.3. Modeling approach

N N

NH

F

OO

HO1

5

4

32

6

7, 8

9

1012

11

19, 20 18

16, 17

1513,14

21

27

26 25

24

2322

Figure 3.6.: Structure of ciprofloxacin. Bonds that are not part of the molecule back-

bone are colored in purble. The 27 relevant bonds are labeled.

In these reaction equations, E symbolizes the excess bonds, that are not covered by

the links in the chain model. For the model, we limit the system to n > b and define

that excess bonds have a higher priority than links when a reaction occurs. If a larger

molecule is broken apart, the excess bonds are distributed to the fragments according

to their share of carbon atoms. The amount of excess bonds for the larger fragment

is defined as β.

β(m) = max

[round

(m bn),b – (n –m – 1)

](3.40)

The applied rounding function is half to even and follows hence the default mode de-

scribed in the IEEE7 Standard for Floating-Point Arithmetic [164]. Through the max

function, it is ensured that fragment exhibits no more than the maximum number of

allowed excess bonds.

The concentration evolution for n ≥ 2 follows the same structure as the two previous

multi-species models (Eqs. 3.36-3.38).

7Institute of Electrical and Electronics Engineers

69

Page 70: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methods

11 10 9 8 7 6 5 4 3 2 1 0

17

16

15

14

13

12

11

10

9

8

7

6

5

4

3

2

1

Number of excess bondsNumberofcarbonatoms

M

A14E7

A15E7

A4E2

A11E5

A15E6

A2E1

A13E6

A3E1

A12E6

A10E5

A5E2

A9E4

A6E3

A7E3

A8E4reaction path examples:

A15E7 A15E6A15E7 A14E7 + M

A15E7 A11E5 + A4E2A15E7 A15–xE7–y + AxEy

Figure 3.7.: Bondmodel reaction schematics for ciprofloxacin with all possible reaction

products of A15E7.

70

Page 71: Modeling and simulation of photocatalytic degradation of ...

3.3. Modeling approach

d

dtCAnEβ(ad) = jn,β,ads – jn,β,des – jn,β,reac + j∗n,β,reac (3.41)

d

dtCAnEβ(aq) = as(jn,β,des – jn,β,ads) (3.42)

d

dtCM(aq) = as j1,0,reac (3.43)

The corresponding fluxes are defined in Eqs. 3.44-3.46.

jn,β,ads = kn,ads (1 – Θ)CAnEβ(aq) (3.44)

jn,β,des = kn,des CAnEβ(ad) (3.45)

jn,β,reac = kn,reac CAnEβ(ad) (3.46)

In Fig. 3.7 all possible fragments for ciprofloxacin and an exemplary reaction pathway

are shown. As in the previous model variant, multiple reaction pathways are possible

for each adsorbed molecule. In addition to the inflow from larger molecules, a new

pathway is added from a fragment with the same size n but with one additional exces-sive bond. The following sum can represent the incoming reaction flux j∗n,b,reac for thissystem.

j∗n,β,reac = nmax∑m=n+1

[2(m–1)–b(m–1)2 km,reac CAm,b(ad)

]+

β+1n–1 kn,reac CAn,β+1(ad) (3.47)

The first part of the equation captures all reactions that yield AnEβ from larger mole-

cules (AnEβ (ad) AmEb (ad) + X). As these reactions have a lower priority, their overallprobability is defined by the ratio of free links (m-1-b) to all chain links (m-1). In the

same way, as described in the fragmentation model, each free link is equally probable

to be broken. Therefore, the same factor of 2/(m – 1) is introduced.

In the second part of the equation, the destruction of an excess bond is taken into

account. Here the probability is (β + 1)/(n – 1). For fragments that are on the left edgeof Fig. 3.7 the probability would be higher than one. A particle with β ≥ n cannot

be formed, as fragments with n = b + 1 can only lose an excess bond. This limitation

71

Page 72: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methodsaddresses the problem of the corner cases as all concentrations where β ≥ n are zero.

3.4. Model implementation

3.4.1. Development objectives

In the previous sections, the theoretical frameworks for the different models were laid

out. As the proposed models cannot be applied to experimental measurements in a

simple way, it is necessary to make an efficient implementation of them available. In

the following paragraphs, this implementation will be discussed.

The goal was to develop a reliable tool that is quickly accessible by experimentalists

as well as reusable for specific cases or new models. To make the software available

on the predominantly used operation systems, Python [165] was chosen as the pro-

gramming language. The quick prototyping capabilities of Python is beneficial during

development. The selection of Python enables an easy route to augment the core

implementation on the fly. Therefore, the user can change the operation of the code

if desired. The toolkit is released under the MIT license8 to facilitate the idea of open

science and incremental improvements through a wider community. This license is

very permissive and requires only the preservation of the copyright and license note.

Derivatives of the software are allowed to be published under different terms in the

future.

While the core library of Python already provides an extensive collection of modules,

it is not feasible to implement the full models in vanilla Python. Therefore, a few ex-

ternal libraries are included that enable practical scientific work under Python. For the

described models, packages from the SciPy ecosystem [166] are used. These are:

• the SciPy core library providing numerical routines,

• NumPy [167] defining new types for arrays and matrices,

• and Matplotlib [168] to render the simulation results.

To prepare the molecule data, scikit-image [169] is used to ascertain the occupied

surface area.

8The full license is available under https://github.com/theia-dev/pdom/blob/master/LICENSE.txt

72

Page 73: Modeling and simulation of photocatalytic degradation of ...

3.4. Model implementationAs the toolkit should be reliable, and external contributions are encouraged, a system

for quality control is needed. For this propose a test-suite was put into practice, to

detect the impact of code changes on the simulation result. As of the time of writing,

16 different tests with 37 individual test-cases are implemented. These tests cover

94% of the core model source code.

The simulations can be started without the user needing to write Python code. This is

archived through the inclusion of two command-line tools. The first one pdom.config

helps to create configuration files. With the configuration file in place, pdom can then

run the simulation and stores the raw results and plots in the file system. Basing the

simulations on configuration files helps to ensure that the results of a simulation can

always be traced back by the parameters that created them.

3.4.2. Molecule parameters

For the simulation, multiple parameters describing the initial molecule are needed.

This data is collected from a central database to simplify the creation of the configu-

ration files. For this project, PubChem was selected. PubChem is an open chemistry

database provided by theNIH9. The required information is collected via an application

programming interface directly from the server.

Two parameters cannot be gathered directly from PubChem for the simulations, the

maximal projected accessible surface, and the excess bond count. The surface area

is calculated based on the 3D structure provided by PubChem. Tools provided by

scikit-image are used to implement the steps described in Sec. 3.3.1 (p. 58). The

backbone bonds in a molecule are counted based on the same structure information

provided by PubChem in an SDF file format. Through the comparison of the backbone

bond count and the number of carbon atoms in the molecule, the excess bond count

is determined.

3.4.3. Solving the differential equation system

The evolution of the concentrations C in the different models is presented as a simpli-fied differential equation.

9National Institutes of Health - USA

73

Page 74: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methods

∂C∂t = f (t,C, · · · ) (3.48)

The right-hand side (RHS) of Eq. 3.48 is implemented based on the different fluxes

described in the previous section. As an example, we can look at the implementation

of the RHS for the single-species model. The function rhs_single10 is called with the

current simulation time t, the number ofmolecules on the surface N[0] and in solution

N[1], and finally the simulation constants k. As none of the fluxes are dependent

on external time coupled fields, the current simulation time t is not utilized in the

calculation of the overall fluxes F.

1 def rhs_single(self, t, N, k):2 ka, kd, kr = k3 # Create empty array for the absolute flux per element4 F = np.zeros(2)5 # Calculate the surface concentration6 C_AS = N[0] / self.cfg['CATALYST']['surface_total']7 # Calculate the volume concentration8 C_A = N[1] / self.cfg['CATALYST']['volume']9 theta = C_AS * self.cfg['MOLECULE']['molar_surface']10 # Calculate the fluxes11 J_ads = C_A * (1.0 - theta) * ka12 J_des = C_AS * kd13 J_reac = C_AS * kr14 F[0] = (-J_des + J_ads - J_reac) * self.cfg['CATALYST']['surface_total']15 F[1] = (+J_des - J_ads) * self.cfg['CATALYST']['surface_total']16 return F

The overall number of molecules is used to ensure that the order of magnitude is

roughly similar for the complete system. This uniformity supports the work of the

solver as the relative, and absolute tolerances can be the same for the whole system.

Furthermore, it simplifies the definition of the overall flux F significantly. As the fluxes

are defined based on the concentration, the molecule count needs to be converted

in lines 6 and 8. The information regarding the surface area or volume is collected in

the central simulation config self.cfg. In line number 9, the coverage of the surface

is determined. After these preparations, the three fluxes can be calculated according

to Eq. 3.11 in lines 11-13. The overall change is then summed up in lines 14 and 15.

10docstring of the original function is omitted

74

Page 75: Modeling and simulation of photocatalytic degradation of ...

3.4. Model implementationThe LSODA subroutine from the FORTRAN ODEPACK library is used to solve these equa-

tions. The LSODA routine is a derivative of the Livermore Solver for Ordinary Differential

Equations (LSODE). It was developed byHindmarsh and Petzold [170, 171] in 1983with

the capability to switch automatically between a nonstiff and stiff solver depending on

the behavior of the problem. For the nonstiff part of a problem, an Adams predictor-

corrector method is used. The stiff parts of a problem are handled by a Backward

Differentiation Formula (BDF) method. The nonstiff method is used initially to solve

the equation system. During the calculation, LSODA can switch dynamically between

the stiff, and the nonstiff method as necessary. For this project, the solve_ivp wrap-

per for LSODA from the SciPy library was used.

3.4.4. Fit to experimental results

Single organic species model

The proposed model cannot be analytically solved in a general way. This restriction

presents a challenge for the fit of experimental data to the models. As the calculation

of the single-species model is rather quick, we can use a custom error function. This

error function allows the selection of different error representations. In the tool are

three different types available: absolute, relative, and relative-squared. The users can

switch between them depending on their datasets.

The minimization workload is handled by Brent’s algorithm for finding a minimum of

a function of one variable [172, 173]. Brent’s method is based on a reliable golden

section search. Depending on the behavior of the function, an inverse parabolic in-

terpolation is used to jump closer to the minima. A different starting interval for the

downhill bracket search is selected for the different available fitting parameters, to

reduce the number of iterations.

Multiple organic species model

The development of the fragments produced by the photocatalytic reaction can be

measured through NPOC or TOC. The concentrations of the segments need to be

summed up to compare the model behavior to the experimental data. The contribu-

tion of a segment An to the total carbon footprint is determined by its carbon-count n.

75

Page 76: Modeling and simulation of photocatalytic degradation of ...

3. Materials and methodsThis results in the following estimation of the TOC:

TOC = 12.0107 gmol–1

NAvo ·∑nmaxn=2 n ·

(NAn(aq) + NAn(ad))V (3.49)

This formula describes the development of the TOC for the whole system. Based on

the experimental setup, it might be appropriate to include only organic fragments in

solution. The user can set this behaviour in the configuration. As this method is an

estimation, it cannot precisely match the initial concentration of the start molecule

with the initial value measured for TOC/NPOC. To enable an efficient fit, the values are

scaled by the initial TOC value. This scaling results in a co-domain from zero to one for

the experimental and simulation data, which is in addition advantageous to the solver.

A damped least-squares method is used to fit the models to the TOC measurements,

specifically, the Levenberg-Marquardt algorithm for its efficiency. While first described

in 1944 by Levenberg [174] it was mostly forgotten, until it was made popular through

the implementation in FORTRAN by Marquardt in 1963 [175]. As the multi-species

models need significantly more computational time compared to the single-species

model, it is essential that the optimization requires as few function evaluations as pos-

sible. SciPy provides a wrapper around the FORTRAN implementation of this algorithm

from MINPACK following Moré’s proposals [176]. An excellent overview regarding the

Levenberg-Marquardt algorithm is provided by Pujol [177].

3.4.5. Availability

The source code for this implementation is available on the CD accompanying this

thesis. Additionally, the same version of the software is archived in the on the open

data platform zenodo under the DOI: 10.5281/zenodo.3953797. The latest source

code is available on GitHub11. Releases of the software are shared on the Python

Package Index12 (PyPI) for easy installation. The handbook belonging to the submitted

version can be found in Appendix - chapter C (p. 105). It describes the usage of the

tool in greater detail and providesmultiple examples. The documentation for the latest

version is available on the Read the Docs platform13.

11https://github.com/theia-dev/pdom12https://pypi.org/project/pdom13https://pdom.readthedocs.org

76

Page 77: Modeling and simulation of photocatalytic degradation of ...

4. Results and discussion

4.1. Nanoparticle properties

The photocatalysts TiO2 and ZnO were characterized to gain an in-depth understand-

ing of their properties and extract relevant parameters for the simulation. For both

powders an SEM image at the same scale is shown in Fig. 4.1. TiO2 particles show a

more delicate structure compared to the ones in the ZnO sample. While the TiO2 pow-

der is composed of monodisperse nanospheres, the ZnO powder presents a more

complex polydisperse rectangular structures.

1.5µm

TiO2 ZnO

10 kV, 5mm, x50000 25 kV, 9mm, x50000

Figure 4.1.: SEM images of colloidal titanium dioxide (TiO2) and zinc oxide (ZnO).

The qualitatively observed differences in the SEM images are directly linked to themea-

77

Page 78: Modeling and simulation of photocatalytic degradation of ...

4. Results and discussionsured specific surface area. This value is derived from BET measurements and used

directly in the calculations. For the P25 TiO2 nanopowder, a value of 56.0m2 g−1 in

agreement with reported data [178] was obtained. In contrast, the surface area of

ZnO was found to be approximately 1/10th at 5.23m2 g−1.

0

20

40

60

80

100

Lightintensity(E365nm

=100) Lamp

TiO2ZnO

436 405 365 313

min

max

Wavelength λ (nm)

Absorbance

Figure 4.2.: Light output of the used mercury fluorescent tubes normalized at 365nm

compared to the diffuse reflectance spectra of TiO2 and ZnO.

The UV-vis diffuse reflectance spectra of ZnO and TiO2 are presented in Fig. 4.2. In ad-

dition, the emission spectrum of the radiation device is provided. It can be observed

that ZnO absorbance already reached its highest level at 365nm. In contrast, the ab-

sorption potential of TiO2 reaches its maxima at higher energies around 310nm. In

general, a smaller difference between the two materials is expected. Due to the small

grain size of the TiO2 powder, its peak is shifted to shorter wavelengths. The lesser

incline of TiO2 is caused by the mixture of different crystal structures in the powder.

The transition between absorbing and non-absorbing is smoothed out by the different

bandgaps. Overall it indicates that the ZnO powder can utilize the light source more

efficiently.

78

Page 79: Modeling and simulation of photocatalytic degradation of ...

4.2. Adsorption-desorption

4.2. Adsorption-desorption

The adsorption rate constants kads were first determined from concentrationmeasure-

ments performed in the dark. The corresponding values, calculated by using Eq. 3.18,

are listed in Tab. 4.1. These values have the same order of magnitude as reported in

[179]. From the asymptotic concentration values measured at adsorption-desorption

equilibrium, CA(aq),∞, the desorption rate constants were determined by the use of

Eqs. 3.19 and 3.20 (cf. Tab. 4.1).

Table 4.1.: Adsorption and desorption rate constants derived from experiments in the

dark for the compounds ciprofloxacin (CIP) and methylene blue (MB) using

photocatalysts TiO2 and ZnO, respectively.

system kads in m s−1 kdes in s−1TiO2-CIP 3.7 • 10−8 2.05 • 10−2

TiO2-MB∗ 2.1 • 10−8 2.85 • 10−2

ZnO-CIP 7.5 • 10−7 9.4 • 10−3

ZnO-MB 5.4 • 10−7 9.3 • 10−3

∗ Fit from degradation experiment

An exception, in the data shown in Tab. 4.1, is the combination of methylene blue with

TiO2. To demonstrate a different approach, the adsorption constant kads was fitted inthis case to data collected under UV light and not in the dark. As shown in the next

section, is the used setup limited by the adsorption speed of the organic molecule.

Therefore, the degradation curve recorded in the experiment is just defined through

the adsorption and desorption constants. In combination with the concentration drop

from the equilibrium phase before the degradation experiment, both parameters can

be determined. On the one hand, this approach makes it possible to reduce the time

investment by eliminating the need for multiple dark experiments. On the other hand,

it is not possible to make a detailed statement about the reaction constant in this case.

The data show that the adsorption rate of methylene blue on ZnO and TiO2 is lower

than that of ciprofloxacin. Comparing the two catalysts, one finds that adsorption on

ZnO is about 20 times faster than on TiO2. This faster adsorption demonstrates that

ZnO presents a more attractive surface for the tested organic molecules compared to

TiO2.

A third approach to extract the adsorption constant kads, is to fit a full time-series of

79

Page 80: Modeling and simulation of photocatalytic degradation of ...

4. Results and discussiondark experiments. In the publication by House et al. [159], their experiments in the

dark are portrayed in rich detail. When the adsorption of a system is as slow as in

this case, determining variations in the concentration close to t = 0 can be challenging.Fitting the experimental data to the single-species model (cf. 3.4.4, p. 75) with the re-

action flux turned off can be used to extract the adsorption and desorption constant

in such a case (Fig. 4.3).

0 0.5 1 1.5 2 2.5 3 3.5 40

5

10

15

20

25

time t (h)

surfa

ceconcentrationC A

(ad)(g/m

2 )

30mg/L 25mg/L15mg/L 5mg/L

Figure 4.3.: Simulation results for the adsoprtion/desorption dynamic of methylene

blue (MB) on TiO2 based on data reported in [159] (as=1.4 • 105m−1). The

marks show the experimental results for the different initial concentra-

tions of MB (5, 15, 25 and 30mg/L). The full line is a fit according to the

single organic species model without reaction flux. The resulting parame-

ters (kads=9.37 • 10−10ms−1, kdes=1.10 • 10−3 s−1) are the same for all four

curves.

While the adsorption constant is over a magnitude smaller kads = 9.78 • 10−10ms−1

compared to the setup used in the experiments presented in this work, the ratio to

the desorption constant kdes = 1.15 • 10−3ms−1 is similar. Differences in the adsorption

value could be due to different pH values of the solutions. For low concentrations (Θ� 1) it follows from Eq. 3.19 that the ratio between the constants is approximately

the same as between the concentration in the solvent and the concentration on the

catalyst surface after the equilibrium is reached (Eq. 4.1).

80

Page 81: Modeling and simulation of photocatalytic degradation of ...

4.3. Photocatalytic degradation

kdeskads ≈CA(aq),∞CA(ad),∞ (4.1)

4.3. Photocatalytic degradation

4.3.1. Single organic species model

The results of the degradation experiments under UV-irradiation are shown in Fig. 4.4.

To fit the measured data points within the framework of the single organic species

model, the adsorption rate constants derived from the experiments conducted in the

dark (Tab. 4.1). The fitting procedure revealed that all values for the reaction rate con-

stant above a certain threshold fit the measured data points. From this observation,

it can be concluded that the adsorption of organic compounds was rate-determining

and that the reaction rate constant is higher than the desorption rate constant (cf. dis-

cussion at the end of Sec. 3.4.4, p. 62). As stated above, the adsorption rate constant

for the system TiO2-methylene blue kads was fitted to the degradation curve, assumingadsorption to be rate-determining.

Further experiments (Fig. 4.5 a) with one half of the TiO2 particle concentration re-

vealed that the degradation of ciprofloxacin slows down. The degradation half-life time

was changed by a factor of 2.05, close to the expected value 2. This corresponds to

themodel prediction in Eq. 3.22 where the rate constant is proportional to the specific

particle surface as. Measurements of the ciprofloxacin concentration for degradationon TiO2 with reduced UV-irradiation are shown in Fig. 4.5 b. Remarkably, the curve

fit yields a reaction rate constant kreac = 0.01 s−1, which is one half of the desorptionrate constant. According to the expression for the apparent rate constant Eq. 3.22,

this means that at this irradiation intensity and below, the degradation rate becomes

reaction-limited.

The slight disagreement between the fitted exponential curve in Fig. 4.5 b and themea-

surement data reflect the limits of the single organic species model. Development

and adsorption of intermediates could modify the degradation curve of ciprofloxacin.

Especially when the reaction speed is the limiting factor, the influence of adsorbed

intermediates plays a more significant role.

81

Page 82: Modeling and simulation of photocatalytic degradation of ...

4. Results and discussion

0

2

4

6

ConcentrationC A

(aq)(102

1m

–3)

simulationexperiment 1experiment 2

0 5 10 15 20 250

2

4

6

Time t (min)0 5 10 15

TiO2 ZnO

ciprofloxacin

methylene

blue

a)

c)

b)

d)

Figure 4.4.: Measured concentration evolutions during the degradation of ciprofloxa-

cin (a,b) and methylene blue (c,d) under UV-irradiation with 1.0 g L−1 cata-

lysts TiO2 (a,c) and ZnO (b,d). The full lines represent fits according to the

above single organic species model using the values of kads and kdes deter-mined from experiments in the dark (Tab. 4.1). The dotted line in panel c)

was obtained by fitting kads from the degradation curve, assuming adsorp-

tion to be rate-determining.

82

Page 83: Modeling and simulation of photocatalytic degradation of ...

4.3. Photocatalytic degradation

0 5 10 15 20 250

2

4

6

8

Time t (min)

ConcentrationC A

(aq)(102

1m

–3)

simulationexperiment 1experiment 2

0 5 10 15 20 25

concentration TiO2 0.5 g L–1 UV intensity 5.5Wm–2

ciprofloxacin

a) b)

Figure 4.5.: Measured concentration evolutions during degradation of ciprofloxacin

with TiO2. Compared to Fig. 4.4 a), a reduced TiO2 particle concentration of

0.5 g L−1 a), and a lower UV-irradiation of 5.5Wm−2 b) was used. The lines

represent fits according to the single organic species model using the val-

ues of kads and kdes determined from the dark experiments (Tab. 4.1). The

fitted reaction constant in panel b) is kreac=0.01 s−1.

4.3.2. Multi organic species model

The concentration evolution of the initial organic compound can be well described by

an exponential decay exp(–kappt) with kapp as the apparent degradation rate constant[159, 161, 162, 179] (cf. Eq. 3.21). Measurements of the evolution of the TOC in the

solution often show an exponential-like decay as well. There are, however, studies

where the TOC does not exhibit such behavior [159, 160, 161, 162]. This mismatch is

likely related to the mineralization kinetics of the intermediate organic compounds. To

demonstrate such complex degradation kinetics of the intermediates, Fig. 4.6 shows

the calculated concentration evolutions of all intermediates based on the the different

variations of the multi organic species model (cf. Sec. 3.4.4, p. 75).

In Figures 4.6 a, c and e, the case of a weak dependence of the desorption rates of

intermediates on the molecule size n is considered (cf. Eq. 3.35). For the incremen-

tal oxidation model (Fig. 4.6 a) as well as for the fragmentation model (Fig. 4.6 c), all

intermediates are mineralized within moderate irradiation time. The excess bonds

model (Fig. 4.6 e) takes the most time of all three models. The smallest intermediate

needs the longest time since its desorption rate is the highest. For the incremental

83

Page 84: Modeling and simulation of photocatalytic degradation of ...

4. Results and discussion

0

1

2

3

4

234

234

0

5

10

15

20

25

234

ConcentrationC A

(aq)(102

1m

–3)

234

0 1 2 3 40

5

10

15

20

25

234

Time t (h)0 1 2 3 4

234

a)

c)

e)

b)

d)

f)

weak strong

increm

ental

fragm

entation

excessbonds

Figure 4.6.: Calculated concentration evolutions of ciprofloxacin intermediates of sizen (curve labels) based on the incremental (a,b), fragmentation (c,d) or ex-

cess bond (e,f) oxidation model with weak (a,c,e) or strong (b,d,f) depen-

dence of the desorption rate constants on the molecule size n accord-

ing to Eqs. 3.33 to 3.35. Parameters: nmax=17, bmax=11, Am=1.028nm2,as=5.6 • 104m−1, kads=3.7 • 10−8ms−1, kdes=2.05 • 10−2 s−1, kreac=0.4 s−1;weak: β1=0.8 s−1; strong: E1=3.0 kJmol−1, E0=44.0 kJmol−1, α1=0.412.

84

Page 85: Modeling and simulation of photocatalytic degradation of ...

4.3. Photocatalytic degradationoxidation model, small intermediates emerge with some delay, whereas in the other

two models, all compounds are produced with similar initial rates. Besides, for the

latter models, particularly high concentrations of small intermediates arise, which are

formed by fragmentation of all larger intermediates, this is contrary to the incremen-

tal oxidation mechanism. As the fragmentation and excess bond models are closely

related, the main difference is the shift of the peaks to later points in time. This shift is

caused by the additional reaction cycles necessary to destroy the excess bonds.

The concentration curves in Figures 4.6 b, d and f show the effect of a strong increase

of the desorption rate of intermediates with decreasing molecule size (cf. Eq. 3.34). An

inhibition of a quick and completemineralization of organic compounds in the solution

is the result. Smaller intermediates remain intact even during long irradiation times.

Due to their high desorption rate their surface coverage is minimal. Possibly, these

small intermediates could be oxidized to a noticeable extent by radicals in the aqueous

solution. This reaction path was, however, neglected in the present model.

The characteristic concentration evolutions in Fig. 4.6 are comparable to experimental

measurements. For example, the plots in Fig. 4.6 c are similar to the evolution of cer-

tain intermediates during photocatalytic degradation of sulfamethoxazole observed

in [37]. In the case of the intensively studied degradation of phenol [180, 181], the

intermediate concentrations show a similar form as well. The evolution of reaction in-

termediates has further been explored for the gas phase oxidation of decane [182].

Many intermediates show concentration evolutions similar to Fig. 4.6 c. However, there

are few crossings of concentration profiles, as seen in Fig. 4.6 b as well.

A detailed view of the excess bond development is presented in Fig. 4.7. This figure

corresponds to the development of the concentration of the intermediates in Fig. 4.6 e.

It shows that while the average number of excess bonds for a given fragment size ndeclines, the overall excess bonds are degraded over the full range of intermediates.

This behavior is a result of the modeling decision to assume that excess bonds are not

specifically targeted.

Measuring the concentration evolution of intermediates directly during an experiment

is quite challenging. To still access information about the intermediates, the overall

amount of organic compounds that remain in the solution can be monitored through-

out the degradation experiment. This value is commonly characterized by measuring

the TOC often in the form of the non-purgable organic carbon (NPOC) (Sec. 3.1.4, p. 49).

85

Page 86: Modeling and simulation of photocatalytic degradation of ...

4. Results and discussion

0 1 2 3 40

2

4

6

8

10

12

17

16

15

Time t (h)

Averagenumbero

fexcessbonds

0

20

40

60

80

100

Overalle

xcessbonds(%)

Figure 4.7.: Change in the overall excess bonds and average number of excess bonds

for intermediates of size n (curve labels). Parameters are choosen as in Fig.4.6 for a weak size dependence of the desorption rate. The curves for the

average number of excess bonds are limited to regions with a concentra-

tions of at least 10m−3 for the corresponding intermediates.

For this reason, we have calculated the evolution of the TOC within this proposedmod-

els. Fig. 4.8 displays the characteristic differences of the TOC evolution between the

three splitting mechanisms discussed above. On the left side, the effect of equalizing

the system in the dark can be perceived. As the adsorption and desorption fluxes are

in equilibrium, the surface is covered with many molecules at the start. Initially, the

degradation process is rate limited by the reaction constants. The intermediates cre-

ated in this phase are still located on the surface. After a few seconds, the surface

concentrations are lowered significantly, so that most reactions are no longer limited

by the reaction rate constant, but through the influx of molecules from the solution.

For the excess bond model, a significant part of the occurring reactions do not con-

tribute to a reduction of the TOC in the beginning but reduce the overall number of

excess bonds. This depletion of reaction potential results in a slower start compared

to the other models. Due to the small-time scale, it is unlikely that this effect can be

captured in an experimental setup.

The right part of Fig. 4.8 is more relevant for comparisons to experimental findings. For

86

Page 87: Modeling and simulation of photocatalytic degradation of ...

4.3. Photocatalytic degradation

0 2 4

3.1

3.2

3.3

3.4Totalorganiccarbon

TOC(mgL

–1)

0 60 120 180 240 0

0.5

1

1.5

2

2.5

3

3.5

Time t (min)

incrementalfragmentexcess bondsweakstrong

Figure 4.8.: Comparison between the TOC evolution for the incremental, the fragmen-

tation, and the excess bonds oxidation model for weak and strong size

dependence of the desorption rate. Parameters are choosen as in Fig. 4.6.

weak size dependence of the desorption rate of intermediates, the incremental oxida-

tion mechanism under moderate irradiation shows an exponential-like decay of the

TOC signal as the organic material disappears. For the fragmentation mechanism and

weak size dependence of the desorption rates of intermediates, the TOC curve shows

an inflection point at the beginning (at about 10min). This slight initial delay in the TOC

signal reduction is caused by the fact that only the smallest fragments are mineralized.

Their formation needs a certain amount of time. Using the excess bonds model, this

inflection point is found later (at around 30min) when the split of a fragment becomes

more likely compared to the removal of an excess bond. At this point in time around

50% of the initial excess bonds are degraded (see Fig. 4.7).

In the case of the incremental degradation mechanism and strong size dependence

of the desorption rate, the TOC signal shows still a fast initial decrease followed by

a considerably slower decay, whereas the fragmentation mechanism shows a more

gradual initial decrease. After a very slow initial phase until around 90min, the degra-

dation rate of the excess bond model increases slightly. An exponential-like tail region

follows this part. For all degradation mechanisms with strong size dependence of the

87

Page 88: Modeling and simulation of photocatalytic degradation of ...

4. Results and discussiondesorption rate, the high desorption constants limit the number of small intermedi-

ates that stay adsorbed on the surface. Therefore, following Eqs. 3.28 and 3.39, these

intermediates will be degradeted slowly. It can be expected that a certain amount of

small organic compounds remains even after a reasonable long irradiation exposure.

0 50 100 150 200 250 300 350 400 02468101214

Time t (min)

Totalo

rganic

carbon

TOC(mg

L–1 )

degradation experimentsimulationTOC experimentfragmentation simulationincremental simulation

010203040506070

Concen

tration

C A(aq

)(µmo

lL–1)

Figure 4.9.: Simulation results for the degradation of methylene blue by TiO2 based

on data reported in [159] (as=1.4 • 105m−1). The full line is a fit ac-

cording to the single organic species model. The resulting parameters

(kads=3.0 • 10−9ms−1, kdes=6.8 • 10−3 s−1) were used for fitting the TOC

curves. For the reaction constant kreac=10 kdes was chosen. The desorp-tion behavior of the intermediates was modeled by Eq. 3.35 with nmax=16(fragmentation model: β0=−6.57 • 10−3 s−1, β1=0.122 s−1; incremental oxi-dation model: β0=−3.06 • 10−2 s−1, β1=0.496 s−1).

Houas et al. [159] studied the degradation of methylene blue on TiO2 and the corre-

sponding TOC evolution. From the presented adsorption experiments, we extracted

the adsorption and desorption rate constants as discussed before (Sec. 4.2, p. 79).

While fitting the degradation curve in preparation with the single-species model, it was

found that the adsorption constant extracted from the experiments in the dark was

to low to explain the presented curve. This analysis suggested that the degradation is

adsorption-limited, i.e. kreac is considerably larger than kdes. Thus, kreac = 10 kreac waschosen in the simulations. The adsorption constants extracted from the degradation

curve under this condition was kads = 3.0 • 10−9ms−1 and while larger than the value

88

Page 89: Modeling and simulation of photocatalytic degradation of ...

4.4. Wastewater treatment plant effluentobtained in the dark it is still in the same order of magnitude. The corresponding des-

orption constant was determined from the equilibrium concentration as shown in Fig.

4.3 as kdes = 6.8 • 10−3 s−1.

Fig. 4.9 shows the simulation results for the TOC data in [159] obtainedwithin the incre-

mental oxidation and fragmentation models and using Eq. 3.35 for the description of

the desorption behavior of intermediates. The number of intermediates was chosen

as the number of carbon atoms inmethylene blue (nmax = 16). Comparison of the fittedTOC curves in Fig. 4.9 with the experimental data shows that both models fit different

stages of the degradation. The long-term evolution is better fitted by the incremental

oxidation model. In the initial phase, the assumption that all connections between in

the abstracted organic molecule components have the same probability of breaking

could be the problem in this case. The delay in the TOC reduction can be explained

with the existence of a preferred breaking point closer to the center of the virtual

molecule. This preference would delay the production of materialized components in

the beginning. Additionally, the approximated adsorption-desorption properties could

be different for specific intermediates. Clearly, for a more detailed quantitative under-

standing of TOC measurements as well as of the evolution of intermediates, a deeper

knowledge of species-specific adsorption-desorption is needed. The incorporation of

a specific reaction tree would prevent the model from being applied universally.

4.4. Wastewater treatment plant effluent

4.4.1. Influence of effluent

The effluent of aWWTP represents a challengingmatrix for photocatalytic degradation.

Due to different parameters like ion concentration, pH value, or the diverse mixture of

organic and inorganicmolecules the concentration development overtimemight differ

from a clean laboratory experiment. To examine this assumption, we selected to trackthe anticonvulsant carbamazepine in a real WWTP effluent sample, alongside other

occurring pharmaceuticals as the ones analyzed in the extended experiment (Tab. 3.2).

Carbamazepine was chosen as a test case, as it was found in a high concentration

(1.29µg L−1) in the effluent.

Two different control experiments were conducted without nanoparticles and without

89

Page 90: Modeling and simulation of photocatalytic degradation of ...

4. Results and discussionUV-irradiation for 45 minutes, to assure that only the combination of nanoparticles

and UV-radiation causes the degradation of the pharmaceuticals. Without nanoparti-

cles, no significant change in the monitored concentrations occurred. After irradiation,

the average change over the monitored drugs was 0.8% (SD 4.8%). However, for the

controls with nanoparticles kept in the dark, we observed an initial drop in the con-

centrations before they stabilized. This condition was reliably reached after the initial

30 minutes. Between the 30 minute mark and the end of the experiment 15 minutes

later the changes in concentration were 0.3% (SD 4.9%) for TiO2 and 0.2% (SD 5.7%)

for ZnO. The initial drop is due to the adsorption on the catalyst’s surface until the

adsorption-desorption equilibrium is reached. These results confirm the idea that the

UV-radiation per se does not induce the degradation process nor the nanoparticles

without UV-irradiation.

The concentration of carbamazepine was, however, considerably reduced over time in

the presence of the catalysts and UV-radiation. Similar behavior was observed for car-

bamazepine present in the wastewater effluent sample as for the ‘clean’ experiment.

In Fig. 4.10 the two experiments with carbamazepine on each of the two catalysts are

compared. It can be noticed that the degradation by ZnO is significantly faster than

with TiO2 in both setups, which is in high accordance with the conducted simulation

of the degradation. These results are shown in Tab. 4.2 and demonstrate that carba-

mazepine can adsorb more than 40 times faster on the ZnO than on the TiO2 surface.

Because of this, ZnO presents a higher degradation rate compared to TiO2 regardless

of its larger surface area.

Table 4.2.: Adsorption and desorption rate constants derived from the experiments

with carbamazepine using photocatalytic TiO2 and ZnO nanoparticles.

catalysts kads in m s−1 kdes in s−1TiO2 5.3 • 10−9 4.5 • 10−4

ZnO 2.2 • 10−7 5.9 • 10−3

When the parameters determined from the Millipore water experiment (Fig. 4.10 a, b)

are compared to the outcome from the effluent (Fig. 4.10 c, d) similar results can be

found. This similarity shows that the effluent background does not reduce the effi-

ciency of the selected photocatalytic particles, and the method to model the degrada-

tion in the Millipore water experiment can be applied for pharmaceuticals investigated

in the effluent as well.

90

Page 91: Modeling and simulation of photocatalytic degradation of ...

4.4. Wastewater treatment plant effluent

020406080100

Degrad

ationra

tioC(t)

/C(0)(%

)

simulationexperiment

0 20 40 60020406080100

Time t (min)0 20 40 60

TiO2 ZnO

isolate

dC(0

)=12.0

mgL–1

effluen

tC(0

)=1.4µ

gL–1

a)

c)

b)

d)

Figure 4.10.: Measured concentration during the degradation of carbamazepine un-

der UV irradiation with 1.0 g L−1 catalysts TiO2 (a,c) and ZnO (b,d). In the

upper row, carbamazepine was dissolved in millipore water (a+b) and

measured by UV-vis spectrophotometry. In the lower row the results

of the degradation of carbamazepine in the treated effluent analysed by

HPLC-MS/MS-method are shown. The full lines represent fits for carba-

mazepine dissolved in millipore water according to the presented model

using the values of kads and kdes determined from experiments in the dark

(Tab. 4.2).

4.4.2. Degradation of pharmaceuticals in the effluent

The degradation results were traced for all target pharmaceuticals and are presented

in the Appendix - chapter B (p. 103). For the SPE extraction, a sample volume of 1mL

was necessary, which allowed a high temporal resolution monitoring with ten samples

throughout one hour. With the larger typical sample size, the experiment would have

been disturbed by taking out to much compared to the initial volume. An overview of

91

Page 92: Modeling and simulation of photocatalytic degradation of ...

4. Results and discussionthe degradation curve based on the two photocatalytic materials is shown in Fig. 4.11.

After 40min, an average degradation of more than 95% for the samples treated with

ZnO was already observed. During the same period, the pharmaceuticals treated with

TiO2 degraded by 40%.

0 10 20 30 40 50 600

20

40

60

80

100

Time t (min)

Degradationratio

C(t)/C(0)(%)

deviationTiO2ZnO

Figure 4.11.: Average degradation ratio over time of the 14 selected pharmaceuticals

(Tab. 3.2) measured by HPLC-MS/MS-method. The grey area corresponds

to the standard deviation.

The simulation shows that treatment with TiO2 would take over four hours to achieve

the same result as ZnO. These differences in degradation rates are even more sig-

nificant, taking into account the smaller surface area of ZnO. The resulting apparent

rate constants and adsorption rate constants are listed in Tab. 4.3 and following the

results from carbamazepine dissolved in Millipore water. In the literature, a variety of

new approaches besides commercial nanoparticles are presented. The apparent rate

constants found, for example from TiO2 nanowires is on average, the same order of

magnitude as the commercial TiO2 particles, but slower compared to the ZnO particles

[183].

Furthermore, a lower selectivity of ZnO compared to TiO2 was observed. In the ZnO

experiments, all selected pharmaceuticals are degraded similarly with just slight devia-

92

Page 93: Modeling and simulation of photocatalytic degradation of ...

4.4. Wastewater treatment plant effluent

Table 4.3.: Apparent rate constants average and adsorption rate constants average for

the 14 studied pharmaceuticals (Tab. 3.2) with the photocatalysts TiO2 and

ZnO.

catalysts kapp in s−1 kads in m s−1

TiO2 1.4 • 10−2 4.0 • 10−9

ZnO 8.6 • 10−2 2.7 • 10−7

tions. This characteristic is significant for applications due to the different mixtures of

pharmaceuticals present in wastewater depending on regional consumption patterns

or seasonal fluctuations. Treating pharmaceuticals present in WWTP effluents using

ZnO and UV-radiation will result in a near-complete degradation of the pollutants in a

relatively short period.

93

Page 94: Modeling and simulation of photocatalytic degradation of ...
Page 95: Modeling and simulation of photocatalytic degradation of ...

5. Conclusions

The simultaneous increase in the worldwide population and living standards in com-

bination with progressive climate change lead to a desperate need for more clean

water. Appropriate technologies for water reuse are an essential part of coping with

this global social challenge. The direct use of treated municipal wastewater for irriga-

tion of agricultural land is at the moment hindered by the presence of organic trace

substances, like pharmaceuticals. When irrigating fertile land, these trace substances

accumulate and subsequently damage soil organisms or are absorbed by plants and

get back into our food chain. Similarly, industrial production processes generate by-

products such as dyes, biocides, and other inert compounds that get into municipal

wastewater or directly into waterways. In contrast to municipal wastewater treatment,

the industrial application is not focused on removing a multitude of low-concentration

substances, but rather the elimination of a few concentrated problematic organic con-

taminants. Photocatalytic degradation is a technology that can potentially resolve

these kinds of challenges. A general model describing the photocatalytic degradation

process is needed to evaluate the performance of differentmaterial combinations and

setups objectively, helping with identification of patterns and optimization pathways.

The creation and evaluation of such a model was the focus of this thesis.

The first set of experiments discussed in this thesis helped to establish the funda-

mental interactions in the model. Through iterations between the model creation

and experimental design, a best-practice setup was implemented to examine the ad-

sorption/desorption equilibrium efficiently. The combination of the model and experi-

ments showed that our setup was limited by the adsorption rate of the test substances

on the catalytic surfaces. By reducing the applied UV-irradiation, we showed that for

the combination of TiO2 and methylene blue, the reaction rate and the adsorption

rate were actually in close approximation to each other. Would these experiments

95

Page 96: Modeling and simulation of photocatalytic degradation of ...

5. Conclusionsonly studied by the common usage of an apparent exponential rate constant, this con-

nection could not be made.

After the general model was established, WWTP effluent was used to evaluate the

model. Compared to the clean experiments carried out to this point, the matrix of

the effluent presents a significant challenge. It was shown that the background did

not hinder the performance of the model, as long as the overall organic load is rea-

sonably low, as to be expected after the ordinary processing steps. The tracked va-

riety of pharmaceuticals could be treated as disconnected cases. As collecting the

adsorption-desorption characteristics of all investigated molecules was not possible

in a reasonable amount of time, the model was used to analyze the complete set as

a virtual average molecule, to compare the performance of the used photocatalytic

material.

A few samples were analyzed using NPOC locally, evaluating the multi-species part

of the model. A systematic investigation of the organic carbon development during

the degradation experiments was not possible in our setup, through the needed sam-

ple size for each measurement point. Therefore, the model was tested successfully

against published data.

Upon entering the thematic field of photocatalytic degradation, it became clear, that

most publications evaluated their experimental results with an exponential fitted ap-

parent rate constant. While this can enable comparison between closely related ex-

periments, it makes it difficult and sometimes even impossible to compare results

between publications. This was one of the reasons to design a model that incorpo-

rates essential parameters describing the experimental system, like the available cat-

alytic surface area, or the approximately area a molecule can cover. Including the

adsorption-desorption dynamics in the model, enables the model to approximate

edge cases, like the linear degradation in the case of high pollutant concentrations.

It allows the estimations of the catalytic surface coverage by the molecules. Besides,

aiding in the evaluation of typical degradation experiments, the multi-species exten-

sion of the model helps to investigate the TOC within the same theoretical framework.

This connects the typically separated experiments and enables a more meaningful in-

vestigation of photocatalytic systems.

While it was tempting to expand the model with more parameters, or a more realistic

reaction network, it is essential to realize that most of these parameters would not be

accessible by experiments. In the end, an over-detailed model would fold back into a

96

Page 97: Modeling and simulation of photocatalytic degradation of ...

system with very similar characteristics to the proposed model, as unknown parame-

ters would need to be merged again. Overall the model is designed to be integrated

into an experimental workflow. To aid the adoption of the model, the incorporation

into existing processes needs to be possible without great hurdles. This was the mo-

tivation to make a ready-to-use implementation of this work available. Providing a

flexible infrastructure facilitates potential extensions to the model, to cover specific

scenarios. For example, multiple target molecules can be simulated at once. Model-

ing a flow reactor-like setup if feasible using an additional reservoir with kreac = 0 and

connect it with fluxes describing the exchange to the reaction reservoir. Moreover,

alterations of the presented models are possible; for example, the alteration of the

target priority of the excess bonds.

While the essential usability of the photocatalytic pathway to remove leftover organic

matter after the conventional wastewater treatment, is well-established in literature

and shown in this work, a quick introduction into the routine treatment process is

unlikely. The biggest hurdles are engineering challenges and the missing knowledge

over the ecological impact of a large scale deployment.

For example, in the WWTP in Dresden-Kaditz a new treatment facility that processes

the water for just 60min would need a water capacity of ≈ 90000 L. In this order

of magnitude, the illumination of the overall system is extremely challenging, as the

UV-light can pass a couple of centimeters through a nanocatalyst solution before it is

mostly absorbed. This problem can be mitigated with the use of a light-guiding sys-

tem made from quartz glass fibers or sheets. Immobilization can be used to avoid

the complex process of separating nanoparticles from a constant water flow. While it

helps to reduce the amount of catalyst material reaching the environment, it severely

reduces the available surface area at the same time. This reduction leads to longer re-

tention times and therefore, an even larger footprint of the potential treatment facility.

Therefore, finding the best catalyst to cut down on the processing time is imperative.

These up-scaling problems can be solved, but they hinder the adoption of this technol-

ogy, especially while other approaches are already successfully employed for the same

task. However, photocatalytic degradation can be an attractive solution, as it is more

energy-efficient than ozone treatment and produces less waste compared to active-

carbon filtration. The most promising path forward towards a large scale application

is probably the synergetic combination with an ozone treatment setup.

While a few challenges remain to be solved, the presentedmodel can help to overcome

97

Page 98: Modeling and simulation of photocatalytic degradation of ...

5. Conclusionsthem. It can assist with the screening of various materials in different configurations

for the use in wastewater treatment setups. By providing comparable results, complex

setups can be optimized in a meaning full way. All in all photocatalytic degradation

should play an important role in a refined wastewater treatment process, and the

presented model can support this effort.

98

Page 99: Modeling and simulation of photocatalytic degradation of ...

Appendix

A. Analytical solution

Solving the system of ordinary differential equations (Eqs. 3.9, 3.10 , p. 58) describing

the single molecule model. This solution is limited to small surface coverage Θ � 1.

The full solution is also presented in Eq. 3.15 and Eq. 3.16 (p. 60).

base system - concentrations

NA(ad) = [CA(ad) · (–kdes – kreac) + CA(aq) · (vads)] · A (A.1)

NA(aq) = [CA(ad) · (kdes) + CA(aq) · (–vads)] · A (A.2)

base system - number of molecules

NA(ad) = NA(ad) · (–kdes – kreac) + NA(aq) · AV(vads) (A.3)

NA(aq) = NA(ad) · (kdes) + NA(aq) · AV(–vads) (A.4)

99

Page 100: Modeling and simulation of photocatalytic degradation of ...

matrix description

~N = P · ~N (A.5)

~N = ( NA(ad)NA(aq)

)(A.6)

θ = A

V(A.7)

P = ( –kdes – kreac vadsθkdes –vadsθ

)(A.8)

find eigenvalues and eigenvectors

a = vadsθ (A.9)

d = kdes (A.10)

r = kreac (A.11)

λ1 = 1

2

(–r – d – a –√r2 + (2d – 2a)r + d2 + 2ad + a2) (A.12)

λ2 = 1

2

(–r – d – a +√r2 + (2d – 2a)r + d2 + 2ad + a2) (A.13)

α = 1

–2d (+r + d – a + √r2 + 2dr – 2ar + d2 + 2ad + a2)–1 (A.14)

β = 1

–2d (+r + d – a – √r2 + 2dr – 2ar + d2 + 2ad + a2)–1 (A.15)

100

Page 101: Modeling and simulation of photocatalytic degradation of ...

α = ( 1

d (λ1 + a)–1)

(A.16)

β = ( 1

d (λ2 + a)–1)

(A.17)

axis transformation

~N = ( α1NA(ad) + α2NA(aq)β1NA(ad) + β2NA(aq)

)(A.18)

~N = ( λ1 0

0 λ2)

~N (A.19)

solution

NA(ad) + d (λ1 + a)–1 NA(aq) = exp(λ1 · t)K1 (A.20)

NA(ad) + d (λ2 + a)–1 NA(aq) = exp(λ2 · t)K2 (A.21)

find constant K with initial conditions

NT + d (λ1 + a)–1 NA(aq) = exp(λ1 · 0)K1 (A.22)

NT + d (λ1 + a)–1 NA(aq) = K1 (A.23)

NT + d (λ2 + a)–1 NA(aq) = exp(λ2 · 0)K2 (A.24)

NT + d (λ2 + a)–1 NA(aq) = K2 (A.25)

101

Page 102: Modeling and simulation of photocatalytic degradation of ...

solve for NA(aq)

exp(λ2 · t)K2 – d (λ2 + a)–1 NA(aq) + d (λ1 + a)–1 NA(aq) = exp(λ1 · t)K1 (A.26)

d [(λ1 + a)–1 – (λ2 + a)–1]NA(aq) = exp(λ1 · t)K1 – exp(λ2 · t)K2 (A.27)

NA(aq) = exp(λ1 · t)K1 – exp(λ2 · t)K2d [(λ1 + a)–1 – (λ2 + a)–1] (A.28)

solve for NA(ad)

NA(aq) = 1

d (exp(λ1 · t)K1 – NA(ad))(λ1 + a) (A.29)

NA(ad) + d (λ2 + a)–1 1d (exp(λ1 · t)K1 – NA(ad))(λ1 + a) = exp(λ2 · t)K2 (A.30)

NA(ad) + (exp(λ1 · t)K1 – NA(ad))λ1 + aλ2 + a = exp(λ2 · t)K2 (A.31)

NA(ad) – NA(ad) λ1 + aλ2 + a = exp(λ2 · t)K2 – exp(λ1 · t)K1 λ1 + aλ2 + a (A.32)

NA(ad)(1 –

λ1 + aλ2 + a) = exp(λ2 · t)K2 – exp(λ1 · t)K1 λ1 + aλ2 + a (A.33)

NA(ad) = exp(λ2 · t)K2 – exp(λ1 · t)K1 λ1 + aλ2 + a1 –

λ1 + aλ2 + a(A.34)

102

Page 103: Modeling and simulation of photocatalytic degradation of ...

B. Effluent pharmaceuticals concentrations

The concentration evolutions for the 14 pharmaceuticals studied in WWTP effluent

(Tab. 3.2, p. 55) are presented in Figures B.1 and B.2. The average degradation ratio is

shown in Fig. 4.11 (p. 92).

0

100

200

300

400

500

ConcentrationC(t)(ngL

–1)

celiprololbezafibratetalinolol

0

100

200

300

400

500 bisoprololvenlafaxine

0 10 20 30 40 50 600

200

400

600

Time t (min)0 10 20 30 40 50 60

eprosartantramadoloxcarbazepin

TiO2 ZnO

a)

c)

e)

b)

d)

f)

Figure B.1.: Concentration evolutions measured by HPLC-MS/MS-method during the

degradation of different pharmaceuticals with low initial concentration un-

der UV-irradiation with 1 g L−1 catalysts TiO2 (left) and ZnO (right). Part 1/2

103

Page 104: Modeling and simulation of photocatalytic degradation of ...

0

0.5

1

1.5

Concen

tration

C(t

)( µgL

−1)

lamotriginecarbamazepinecandesartanirbesartan

0

1

2

3 valsartan

0 10 20 30 40 50 60024681012

Time t (min)0 10 20 30 40 50 60

gabapentin

TiO2 ZnO

a)

c)

e)

b)

d)

f)

Figure B.2.: Continued from Fig. B.1. Concentration evolutions measured by HPLC-

MS/MS-method during the degradation of different pharmaceuticals with

low initial concentration under UV-irradiation with 1 g L−1 catalysts TiO2

(left) and ZnO (right). Part 2/2

104

Page 105: Modeling and simulation of photocatalytic degradation of ...

pdom Handbook

Release 1.0.0

simulation toolkit for the Photocatalytic Degradation of Organic Molecules

Contents

1 Setup 11.1 CLI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Library . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.3 Developing and documentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.4 pipx . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

2 Usage 32.1 CLI pdom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32.2 CLI pdom.config . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32.3 Library . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

3 Configuration 43.1 Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43.2 Experimental data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53.3 Configuration settings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

4 Examples 104.1 Adsorption - Desorption equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104.2 Degradation fit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124.3 TOC simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154.4 TOC fit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

5 Source documentation 245.1 pdom.Simulate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245.2 pdom.data.Parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265.3 pdom.Molecule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295.4 pdom.export . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

References 33

1

C. pdom handbook

105

Page 106: Modeling and simulation of photocatalytic degradation of ...

1 Setup

1.1 CLI

Most features of pdom can be accessed directly from its command-line tools. The easiest way to install them onyour system is via pipx. pipx is a packagemenager for tools written in python that helps to keep them isolated andup to date. Some notes on how to get pipx running can be found at the end of this section. If you just need theCLI, use the following line to install pdom.

$ pipx install pdom

1.2 Library

You can also install the pdom library directly through the Python Package Index (PyPI) for use in your ownprojects. The use of a virtual environment is recommended.

$ pip install pdom

If the stable version of pdom on PyPI is missing a particular function, you can install the latest developmentversion directly from the GitHub repository.

$ pip install -U git+https://github.com/theia-dev/pdom.git#egg=pdom

1.3 Developing and documentation

To work with the source code clone the repository from GitHub and install the requirements. The source code isaccompanied by the documentation and an extensive collection of test cases.

$ git clone https://github.com/theia-dev/pdom.git$ python3 -m venv pdom_env$ . pdom_env/bin/activate(pdom_env) $ pip install --upgrade pip(pdom_env) $ pip install -r pdom/requirements.txt

Building the documentation locally needs a few extra python packages.

(pdom_env) $ pip install -r pdom/docs/requirements.txt

The HTML version of the documentation can then be built:

(pdom_env) $ sphinx-build -b html pdom/docs pdom/docs_html

The tests are located under pdom/tests can be started through:

(pdom_env) $ python -m unittest discover -s pdom/tests

2

106

Page 107: Modeling and simulation of photocatalytic degradation of ...

1.4 pipx

Under macOS:

For macOS the Homebrew package manager is the easiest way to install pipx.

$ brew install pipx$ pipx ensurepath

Under Linux:

For some distributions the python package system pip is not installed by default. On Debin/Ubuntu systems it canbe quickly installed.

$ sudo apt update$ sudo apt install python3-pip

Then pipx can be added.

$ python3 -m pip install --user pipx$ python3 -m pipx ensurepath

Under Windows:

Python is not installed by default under Windows. You can get the installer from the Python download page. Thepython package system pip is already included in the latest releases.

In the windows commandline pipx can then be installed.

$ python3 -m pip install --user pipx$ python3 -m pipx ensurepath

For more information on pipx refer to its documentation documentation.

3

107

Page 108: Modeling and simulation of photocatalytic degradation of ...

2 Usage

Here you can find a quick summary about pdom command-line tools and the pdom library. A more in-depthoverview can be found in the Examples section

2.1 CLI pdom

A simple command line interface to run pdom.

usage: pdom [-h] [-d DATA] [-o OUT] config [config ...]

Positional Arguments

config config file (.ini)

Named Arguments

-h, --help show this help message and exit

-d, --data experimental data for fit (.json)

-o, --out output folder (absolute or relative, will be created if it does not exit)

2.2 CLI pdom.config

A command line interface to create pdom config files.

usage: pdom.config [-h] [-s STRUCTURE] [-o OUTFILE]

Named Arguments

-h, --help show this help message and exit

-s, --structure xyz molecule structure file

-o, --outfile output file

How to create a config is described in the Configuration section.

2.3 Library

import pdom

simulation = Simulate('simple.ini')simulation.run()

See also the source code documentation of pdom.Simulate.

4

108

Page 109: Modeling and simulation of photocatalytic degradation of ...

3 Configuration

3.1 Simulation

To run the simulation pdom needs a couple of information from the user. For pdom, this data is saved in an .inifile. This ensures that the results of a simulation can always be accompanied by the parameters that created them.

A simple config for pdom looks like this:

Listing 1: simple.ini

[SIMULATION]id = simple_degradation_methylene_bluemulti = Falsefit = Falseduration = 5 h

[SYSTEM]concentration_solution = 0.08 mmol/Lk_ads = 9.0E-10 m/sk_des = 1.0E-3 1/sk_reac = 1.1E-2 1/s

[CATALYST]concentration = 2.0 g/Lsurface = 50 m^2/gvolume = 1e-3 m^3

[MOLECULE]name = methylene bluecomposition = C16H18S1N3molar_volume = 226.6 Ang^3/moleculemolar_surface = 99.7 Ang^2/moleculediffusion_model = s

The parameters are arranged in different sections. Not all sections need to exist for each simulation.

To create a new .ini file you can use the configuration tool of pdom.

$ pdom.config --out 'my_new.ini'

It will guide you through the process by collecting all relevant information. For the parameters of the initialmolecule, it is helpful to look up its chemID in PubChem before you start the process. This enables the automaticgathering of the molecule parameters from this database.

Most of the time, the automatic process should suffice. But if you want to build your .ini file from scratch, takea look into the available Configuration settings.

5

109

Page 110: Modeling and simulation of photocatalytic degradation of ...

3.2 Experimental data

To extract parameters, we need to compare experimental data to the simulation. This data needs to be providedin a structured way. For pdom we use a .json file. Depending on the type of fit you want to carry out, theavailable features differ slightly.

Adsorption-Desorption experiments

Adsorption-Desorption experiments in the dark are analyzed with the single-species model. As it is common tohave multiple repetitions that are based on the same setup. pdom supports multiple time series in its fits. Theinitial concentration and time steps can be different between the series. Below are examples of such a data file.

Listing 2: ads_des_multi.json

{"time_series": [

[[0, 10, 15, 30, 60, 120, 240],[0.000, 0.076, 0.143, 0.175, 0.199, 0.207, 0.209]

], [[0, 10, 15, 30, 60, 120, 240],[0.000, 0.251, 0.452, 0.570, 0.609, 0.637, 0.633]

], [[0, 10, 15, 30, 60, 120, 240],[0.000, 0.598, 0.750, 0.898, 0.998, 1.021, 1.020]

], [[0, 10, 15, 30, 60, 120, 240],[0.000, 0.700, 0.996, 1.248, 1.396, 1.415, 1.415]

]],"time_series_meta": [

{"unit": "min","type": "t"

},{

"unit": "mo/mc","factor": 1E-3,"type": "surface"

}],"initial_concentration": [5.0, 15.0, 25.0, 35.0],"initial_concentration_meta": {

"unit": "mg/L","type": "solution"

}}

6

110

Page 111: Modeling and simulation of photocatalytic degradation of ...

Degradation experiments

For degradation experiments all time series have to start with the same initial concentration set in the .ini file.This is the data file from the example Degradation fit.

Listing 3: fit_reac_mutli.json

{"time_series": [

[[0, 1, 2, 3, 4, 5, 6, 7, 8, 10, 12, 14, 16, 20, 25],[3.442, 2.847, 2.428, 2.229, 1.949, 1.801, 1.705, 1.535,1.315, 1.021, 0.9690, 0.8238, 0.5114, 0.2839, 0.1353]

], [[0, 1, 2, 3, 4, 5, 6, 7, 8, 10, 12, 14, 16, 20, 25],[3.505, 2.876, 2.586, 2.323, 2.242, 2.097, 1.781, 1.681,1.461, 1.249, 0.9543, 1.0852, 0.7665, 0.4324, 0.2368]

]],"time_series_meta": [

{"unit": "min","type": "t"

}, {"unit": "mg/L","type": "solution"

}]

}

Multi species model

To compare multi-species model simulations to experiments, TOC (or NPOC) can be used. In general, the fit toTOC data is the last step in the experiment analytics. This fit is limited to a single TOC curve, due to the usuallydemanding experimental process. If multiple TOC experiments are available, the results should be average beforesimulation. As just one initial concentration is needed, it is taken from the .ini file.

7

111

Page 112: Modeling and simulation of photocatalytic degradation of ...

This is the data file from the example TOC fit.

Listing 4: fit_toc.json

{"time_series": [

[0, 60, 120, 180, 360],[12.6, 8.8, 6.78, 4.1, 2.77]

],"time_series_meta": [

{"unit": "min","type": "t"

}, {"unit": "mg/L","type": "toc"

}]

}

3.3 Configuration settings

SIMULATION

• id → str

default example_mb

• multi → bool

default True

note MULTI section needed if True

• fit → bool

default False

note FIT section needed if True

• duration → float

default 5 h

units h, min, s

SOLVER

Relative and absolute tolerances for the LSODA solver

• rtol → float

default 1e-9

• atol → float

default 1e-5

ENVIRONMENT

• temperature → float

8

112

Page 113: Modeling and simulation of photocatalytic degradation of ...

default 20 C

units K, C

CATALYST

• concentration → float

default 2.5 g/L

units g/m^3, g/L, mg/L

• surface → float

default 56e3 m^2/g

units m^2/g, cm^2/g

• volume → float

default 1e-3 m^3

units m^3, L, cm^3, mL

MOLECULE

• name → str

default methylene blue

• composition → chem_formula

default C16H18S1N3

• excess_bonds → int

default 14

• molar_volume → float

default 226.6 Ang^3/molecule

units Ang^3/molecule, nm^3/molecule, cm^3/mol

• molar_surface → float

default 99.7 Ang^2/molecule

units Ang^2/molecule, nm^2/molecule, m^2/molecule

• diffusion_model → str

default s

note s: Stokes (default), wc: Wilke-Chang, hm: Hayduk-Minhas

SYSTEM

• concentration_solution → float

default 10 mg/L

units molecule/m^3, molecule/L, mol/m^3, mmol/L, M, mol/L, mo/mc, g/L,mg/L, g/m^3

• concentration_surface → float

default 0

9

113

Page 114: Modeling and simulation of photocatalytic degradation of ...

units molecule/m^2, mol/m^2, g/m^2, mg/m^2

note if concentration_surface is not set system is considered in equilibrium (dark)

• k_ads → float

default 1E-9 m/s

unit m/s

• k_des → float

default 1E-3 1/s

unit 1/s

• k_reac → float

default 1E-2 1/s

unit 1/s

FIT

This section is just active if fit is True

• type → str

default dark

note dark, reaction or toc

• search → str

default relative

note minima search absolute, relative or relative_square

note does not apply to fit type toc

• max_step → int

default 100

note does not apply to fit type toc

MULTI

This section is just active if multi is True

• split_model → str

default fragmentation

note incremental, fragmentation or excess_bonds

• desorption_model → str

default weak

note weak or strong

• TOC_estimation → str

default all

note all or volume

• segment_export → str

10

114

Page 115: Modeling and simulation of photocatalytic degradation of ...

default mass

note mass or molecule_count

MULTI_WEAK

This section is just active if desorption_model is set to weak. Just one value needed if k_des isset.

• beta_0 → float

default -0.029 1/s

unit 1/s

• beta_1 → float

default 0.8 1/s

unit 1/s

MULTI_STRONG

This section is just active if desorption_model is set to strong

• E_0 → float

default 44.0 kJ/mol

unit kJ/mol

• E_1 → float

default 3.0 kJ/mol

unit kJ/mol

• alpha_0 → float

default 1.51e8 1/s

unit 1/s

• alpha_1 → float

default 0.412

4 Examples

4.1 Adsorption - Desorption equilibrium

In this first example, the goal is to simulate a system reaching equilibrium in the dark. To run the simulation, weare going to create a configuration file for pdom first. After calling pdom.config we need to answer a fewquestions. The parts which require user input are highlighted in grey.

$ pdom.configID of the system (avoid spaces): example_ads_des

Should data be fitted to the simulation?1: fit2: just simulation

(continues on next page)

11

115

Page 116: Modeling and simulation of photocatalytic degradation of ...

(continued from previous page)

Your choice: 2

What kind of simulation?1: Adsorption-Desorption2: Degradation3: TOC

Your choice: 1

How can you identify the initial molecule?1: chemID (https://pubchem.ncbi.nlm.nih.gov)2: name

Your choice: 1Molecule: 2764Found ciprofloxacin (C17H18FN3O3)

What is the catalyst concentration?the allowed unis are: g/m^3, g/L, mg/L

Value: 1 g/L

What is the catalyst surface area?the allowed unis are: m^2/g, cm^2/g

Value: 56 m^2/g

What is the overall volume?the allowed unis are: m^3, L, cm^3, mL

Value: 0.3 L

How long should the simulation be?the allowed unis are: h, min, s

Value: 15 min

What is the adsorption constant?the allowed unis are: m/s

Value: 3.7E-8 m/s

What is the desorption constant?the allowed unis are: 1/s

Value: 2.0E-2 1/s

What is concentration in the solution?the allowed unis are: molecule/m^3, molecule/L, mol/m^3, mmol/L, M, mol/L, mo/

→˓mc, g/L, mg/L, g/m^3Value: 0.1 mmol/L

What is concentration on the surface?the allowed unis are: molecule/m^2, mol/m^2, g/m^2, mg/m^2

Value: 0 mol/m^2

12

116

Page 117: Modeling and simulation of photocatalytic degradation of ...

The resulting file example_ads_des.ini is now in your working directory. To start the simulation simplycall pdom:

$ pdom example_ads_des.iniStart calculating single species modelCalculation finished!Results saved in <your_working_dir>/example_ads_des

In the newly created folder <your_working_dir>/example_ads_des, you find the raw data files withcorresponding units and a plot of the concentration development over time.

0 2 4 6 8 10 12 14time t (min)

0

5

10

15

20

25

30

volu

me

conc

entr

atio

n C

(g/m

3 )

0

10

20

30

40

50

surf

ace

conc

entr

atio

n C

(g/

m2 )

Fig. 1: Development of the concentration in solution and on the surface during the equilibration in the dark.

4.2 Degradation fit

Using the single species model, we fit in this example the reaction constant 𝑘reac to experimental data. Withpdom.config we can create example_reac_fit.ini. The parts which require user input are highlightedin grey.

$ pdom.configID of the system (avoid spaces): example_reac_fit

Should data be fitted to the simulation?1: fit2: just simulation

Your choice: 1

What kind of experiment was conducted?(continues on next page)

13

117

Page 118: Modeling and simulation of photocatalytic degradation of ...

(continued from previous page)

1: Adsorption-Desorption2: Degradation3: TOC

Your choice: 2

How can you identify the initial molecule?1: chemID (https://pubchem.ncbi.nlm.nih.gov)2: name

Your choice: 1Molecule: 2764Found ciprofloxacin (C17H18FN3O3)

What is the catalyst concentration?the allowed unis are: g/m^3, g/L, mg/L

Value: 1.0 g/L

What is the catalyst surface area?the allowed unis are: m^2/g, cm^2/g

Value: 56 m^2/g

What is the overall volume?the allowed unis are: m^3, L, cm^3, mL

Value: 1 L

How long should the simulation be?the allowed unis are: h, min, s

Value: 0.8 h

Which constant is known?1: k_ads2: k_des

Your choice: 1

What is the adsorption constant?the allowed unis are: m/s

Value: 3.7E-8 m/s

What is concentration in the solution?the allowed unis are: molecule/m^3, molecule/L, mol/m^3, mmol/L, M, mol/L, mo/

→˓mc, g/L, mg/L, g/m^3Value: 3.8 mg/L

14

118

Page 119: Modeling and simulation of photocatalytic degradation of ...

The resulting file example_reac_fit.ini is now in your working directory. Next, the experimental dataneeds to be stored as example_reac_fit.json. In this example, two time series with the same initialconcentrations are used.

Listing 5: example_reac_fit.json

{"time_series": [

[[0, 1, 2, 3, 4, 5, 6, 7, 8, 10, 12, 14, 16, 20, 25],[3.442, 2.847, 2.428, 2.229, 1.949, 1.801, 1.705, 1.535,1.315, 1.021, 0.9690, 0.8238, 0.5114, 0.2839, 0.1353]

], [[0, 1, 2, 3, 4, 5, 6, 7, 8, 10, 12, 14, 16, 20, 25],[3.505, 2.876, 2.586, 2.323, 2.242, 2.097, 1.781, 1.681,1.461, 1.249, 0.9543, 1.0852, 0.7665, 0.4324, 0.2368]

]],"time_series_meta": [

{"unit": "min","type": "t"

}, {"unit": "mg/L","type": "solution"

}]

}

With both files prepared pdom can be started.

$ pdom example_reac_fit.ini --data example_reac_fit.jsonStart fitting to data from reaction experiment.Fit finished after 24 iterations.

k_ads: 3.700E-08 m/sk_des: 2.180E-02 1/sk_reac: 1.452E-01 1/serror: 3.136E+00

Results saved in <your_working_dir>/example_reac_fit

The result of the fit is stored under <your_working_dir>/example_reac_fit/fit_single.json.

15

119

Page 120: Modeling and simulation of photocatalytic degradation of ...

0 10 20 30 40 50time t (min)

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

volu

me

conc

entr

atio

n C

(g/m

3 )

EXP 00EXP 01

Fig. 2: Concentration measurement of a degradation experiment compared with the simulated development.

Listing 6: <your_working_dir>/example_fit_reac/fit_single.json

{"k_ads": "3.700E-08 m/s","k_des": "2.180E-02 1/s","k_reac": "1.452E-01 1/s","error": "3.136E+00","iterations": 19

}

In the same folder, you find the raw data files with corresponding units. The saved plot shows the concentrationdevelopment over time compared to the experimental results.

4.3 TOC simulation

Fragmentation

When the overall degradation of an organic molecule should be simulated, one of the multi-species models mustbe used. Three different split models are available in pdom:

• incremental

• fragmentation

• excess bonds

16

120

Page 121: Modeling and simulation of photocatalytic degradation of ...

To model the generalized split species, the size dependents of the desorption constant 𝑘ads must be described. Youcan either choose strong or weak dependents for your simulation. We chose weak dependents in this example,due to the fewer needed parameters.

As before the config file example_toc_sim.ini can be generated with pdom.config. Lines with requireuser input are highlighted in grey.

$ pdom.configID of the system (avoid spaces): example_toc_sim

Should data be fitted to the simulation?1: fit2: just simulation

Your choice: 2

What kind of simulation?1: Adsorption-Desorption2: Degradation3: TOC

Your choice: 3

How can you identify the initial molecule?1: chemID (https://pubchem.ncbi.nlm.nih.gov)2: name

Your choice: 1Molecule: 4139Found methylene blue cation (C16H18N3S+)

What is the catalyst concentration?the allowed unis are: g/m^3, g/L, mg/L

Value: 1 g/L

What is the catalyst surface area?the allowed unis are: m^2/g, cm^2/g

Value: 56 m^2/g

What is the overall volume?the allowed unis are: m^3, L, cm^3, mL

Value: 0.4 L

How long should the simulation be?the allowed unis are: h, min, s

Value: 4 h

Which split model should be used?1: incremental2: fragmentation3: excess_bonds (slow)

Your choice: 2

Is the system in equilibrium (dark)?1: Yes2: No

Your choice: 1

(continues on next page)

17

121

Page 122: Modeling and simulation of photocatalytic degradation of ...

(continued from previous page)

Which model should be usedfor the estimation of size dependent k_des?

1: Strong2: Weak

Your choice: 2

What is the adsorption constant?the allowed unis are: m/s

Value: 2.1E-8 m/s

What is the desorption constant?the allowed unis are: 1/s

Value: 2.85E-2 1/s

What is the reaction constant?the allowed unis are: 1/s

Value: 2.85E-1 1/s

What is the value of beta_1?the allowed unis are: 1/s

Value: 0.2 1/s

What is concentration in the solution?the allowed unis are: molecule/m^3, molecule/L, mol/m^3, mmol/L, M, mol/L, mo/

→˓mc, g/L, mg/L, g/m^3Value: 4 mg/L

With the generated config example_toc_sim.ini we can start the simulation with pdom.

$ pdom example_toc_sim.iniStart calculating multi species modelCalculation finished!Results saved in <your_working_dir>/example_toc_sim

In the folder <your_working_dir>/example_toc_sim, you find the raw data files with correspondingunits and three plots. The first plot is an overview with TOC and the concentration of the initial molecule insolution. The other two show concentration developments of the segments in solution and on the surface.

18

122

Page 123: Modeling and simulation of photocatalytic degradation of ...

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0time t (h)

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

volu

me

conc

entr

atio

n C

(g/m

3 )

0.0

0.5

1.0

1.5

2.0

2.5

TOC C

(g/m

3 )

Fig. 3: Development of the concentration in solution and the total organic carbon (TOC) using the fragmentationsplit model.

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0time t (h)

10 6

10 5

10 4

10 3

10 2

10 1

volu

me

conc

entr

atio

n C

(g/m

3 )

2

4

6

8

10

12

14

carb

on c

ount

Fig. 4: Development of the segments in solution when using the fragmentation split model.

19

123

Page 124: Modeling and simulation of photocatalytic degradation of ...

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0time t (h)

10 8

10 7

10 6

10 5

10 4

10 3

10 2

10 1

surf

ace

conc

entr

atio

n C

(g/

m2 )

2

4

6

8

10

12

14

carb

on c

ount

Fig. 5: Development of the segments on the surface when using the fragmentation split model.

Incremental

Because pdom is based on configuration files, you can simply make a copy of example_toc_sim.ini andchange the settings quickly for a new simulation. You can alter the split_model key in the section MULTI toincremental for example:

...[MULTI]split_model = incrementaldesorption_model = weak...

This generates the following results:

20

124

Page 125: Modeling and simulation of photocatalytic degradation of ...

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0time t (h)

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

volu

me

conc

entr

atio

n C

(g/m

3 )

0.0

0.5

1.0

1.5

2.0

2.5

TOC C

(g/m

3 )

Fig. 6: Development of the concentration in solution and the total organic carbon (TOC) using incremental splitmodel.

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0time t (h)

10 7

10 6

10 5

10 4

10 3

10 2

10 1

100

volu

me

conc

entr

atio

n C

(g/m

3 )

2

4

6

8

10

12

14

carb

on c

ount

Fig. 7: Development of the segments in solution when using the incremental split model.

21

125

Page 126: Modeling and simulation of photocatalytic degradation of ...

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0time t (h)

10 7

10 6

10 5

10 4

10 3

10 2

10 1

surf

ace

conc

entr

atio

n C

(g/

m2 )

2

4

6

8

10

12

14

carb

on c

ount

Fig. 8: Development of the segments on the surface when using the incremental split model.

4.4 TOC fit

To fit TOC data with pdom, one of the multi-species models must be selected:

• incremental

• fragmentation

• excess bonds

For this example, we will use incremental. The model for the desorption constant 𝑘ads is always weak when aTOC experiment is fitted.

The generation of the config file example_toc_fit.ini is again carried out with pdom.config. Lineswith require user input are highlighted in grey.

$ pdom.configID of the system (avoid spaces): example_toc_fit

Should data be fitted to the simulation?1: fit2: just simulation

Your choice: 1

What kind of experiment was conducted?1: Adsorption-Desorption2: Degradation3: TOC

Your choice: 3(continues on next page)

22

126

Page 127: Modeling and simulation of photocatalytic degradation of ...

(continued from previous page)

How can you identify the initial molecule?1: chemID (https://pubchem.ncbi.nlm.nih.gov)2: name

Your choice: 1Molecule: 4139Found methylene blue cation (C16H18N3S+)

What is the catalyst concentration?the allowed unis are: g/m^3, g/L, mg/L

Value: 2.5 g/L

What is the catalyst surface area?the allowed unis are: m^2/g, cm^2/g

Value: 56 m^2/g

What is the overall volume?the allowed unis are: m^3, L, cm^3, mL

Value: 1 L

How long should the simulation be?the allowed unis are: h, min, s

Value: 6 h

Which split model should be used?1: incremental2: fragmentation3: excess_bonds (slow)

Your choice: 1

Is the system in equilibrium (dark)?1: Yes2: No

Your choice: 1

Which parameter should be fitted?1: k_reac2: beta_1

Your choice: 2

What is the adsorption constant?the allowed unis are: m/s

Value: 3.0E-9 m/s

What is the desorption constant?the allowed unis are: 1/s

Value: 6.8E-3 1/s

What is the reaction constant?the allowed unis are: 1/s

Value: 6.8E-2 1/s

What is concentration in the solution?

(continues on next page)

23

127

Page 128: Modeling and simulation of photocatalytic degradation of ...

(continued from previous page)

the allowed unis are: molecule/m^3, molecule/L, mol/m^3, mmol/L, M, mol/L, mo/→˓mc, g/L, mg/L, g/m^3Value: 0.069 mmol/L

After the config is generated, the experimental data set is created. In this example, values published by Houas(2001) [Hou01] will be used.

Listing 7: example_reac_fit.json

{"time_series": [

[0, 60, 120, 180, 360],[12.6, 8.8, 6.78, 4.1, 2.77]

],"time_series_meta": [

{"unit": "min","type": "t"

}, {"unit": "mg/L","type": "toc"

}]

}

With both files prepared, pdom can be started.

$ pdom example_toc_fit.ini --data example_toc_fit.jsonStart fitting to toc

Iteration Total nfev Cost Cost reduction Step norm→˓Optimality

0 1 2.0681e-02 3.→˓05e+00

1 2 4.6195e-03 1.61e-02 1.00e-01 6.→˓13e-01

2 3 2.7331e-03 1.89e-03 6.05e-02 5.→˓52e-02

3 4 2.7079e-03 2.52e-05 8.61e-03 8.→˓37e-04

4 5 2.7079e-03 1.79e-08 1.39e-04 1.→˓54e-04

5 6 2.7079e-03 1.99e-09 2.55e-05 5.→˓45e-04

`xtol` termination condition is satisfied.Function evaluations 6, initial cost 2.0681e-02, final cost 2.7079e-03, first-→˓order optimality 5.45e-04.Fit finished

k_ads: 3.000E-09 m/sk_des: 6.800E-03 1/sk_reac: 6.800E-02 1/sbeta_0: -1.003E-02 1/sbeta_1: 2.693E-01 1/serror: 4.657E-02

Results saved in <your_working_dir>/example_toc_fit

24

128

Page 129: Modeling and simulation of photocatalytic degradation of ...

The result of the fit is stored under <your_working_dir>/example_toc_fit/fit_toc.json.

Listing 8: <your_working_dir>/example_toc_fit/fit_toc.json

{"k_ads": "3.000E-09 m/s","k_des": "6.800E-03 1/s","k_reac": "6.800E-02 1/s","beta_0": "-1.003E-02 1/s","beta_1": "2.693E-01 1/s","sd_error": "4.657E-02"

}

In the same folder, you find the raw data files with corresponding units. The saved plot shows the TOC devel-opment over time compared to the experimental results.

0 1 2 3 4 5 6time t (h)

0

2

4

6

8

10

12

TOC

(g/m

3 )

simulationexperiment

Fig. 9: Experimental TOC measurements compared with the simulation results.

5 Source documentation

5.1 pdom.Simulate

class pdom.Simulate(config_file, out_folder=None, data_file=None, overwrites=None, resolu-tion=250)

Class to simulation different degradation models

Parameters

• config_file (str, Path) – .ini file to load

25

129

Page 130: Modeling and simulation of photocatalytic degradation of ...

• out_folder (str, Path, optional) – folder to save the results

• data_file (str, Path, optional) – .json file containing experimental data

• overwrites (dict, optional) – overwrite settings from the config file

• resolution (int, optional) – time resolution for data export

cfgthe simulation configuration created with pdom.data.Parameter from the config file.

export_ttime steps used for export - filled after run()

rhs_bonds(t, N_flat, k, N_shape)Right-hand site for multi species model excess_bonds.

Calculate the derivative for a given concentration profile.

Parameters

• N_flat (ndarray) – number of molecules | 3-dim flattened

• t (float) – time (not used)

• k (ndarray) – simulation constants (𝑘ads, 𝑘des, 𝑘reac)

• N_shape (tuple) – original shape of N

• first dimension of N corresponds to the number of Carbon atoms carbon_count=index+1

• second dimension of N corresponds to the number of excess bonds excess_bonds=index

• max(excess_bonds) < max(carbon_count)

• N[:, :, 0] is the number of molecules on the surface

• N[:, :, 1] is the number of molecules in solution

• the simulation constants must have the shape (3, max(excess_bonds)+1,max(carbon_count))

Returns first derivative

Return type ndarray

rhs_multi(t, N_flat, k, N_shape)Right-hand site for multi species models fragmentation and incremental.

Calculate the derivative for a given concentration profile.

Parameters

• N_flat (ndarray) – number of molecules | 2-dim flattened

• t (float) – time (not used)

• k (ndarray) – simulation constants (𝑘ads, 𝑘des, 𝑘reac)

• N_shape (tuple) – original shape of N

• first dimension of N corresponds to the number of Carbon atoms carbon_count=index+1

• N[:, 0] is the number of molecules on the surface

26

130

Page 131: Modeling and simulation of photocatalytic degradation of ...

• N[:, 1] is the number of molecules in solution

• the simulation constants must have the shape (3, max(carbon_count))

Returns first derivative

Return type ndarray

rhs_single(t, N, k)Right-hand site for single species model.

Calculate the derivative for a given concentration profile.

Parameters

• N (ndarray) – number of molecules | 1-dim

• t (float) – time (not used)

• k (ndarray) – simulation constants (𝑘ads, 𝑘des, 𝑘reac)

• N[0] is the number of molecules on the surface

• N[1] is the number of molecules in solution

Returns first derivative

Return type ndarray

run()Runs either a simulation or performs a parameter fit based on the information stored in cfg. Theresults are saved to the file system.

ttime steps used by the solver - filled after run()

5.2 pdom.data.Parameter

class pdom.Parameter(config_file, data_file=None, overwrites=None)Collection of constants, parameters and helper functions.

This class loads the configuration and handles experimental data. Provide functions to convertdata into pdom base units and alter settings.

Parameters

• config_file (str, Path) – .ini file to load

• data_file (str, Path, optional) – .json file containing experimental data

• overwrites (dict, optional) – overwrite settings from the config file

avogadro = 6.02214129e+23𝑁𝐴 - Avogadro constant [1/mol]

static get_diffusion_constant_air(mol_volume, molar_weight, temperature=293.15,pressure=101325.0)

Diffusion constant in air depending on molar volume and weight Based on the FSG model with con-stants from the Handbook of chemical property estimation methods.

27

131

Page 132: Modeling and simulation of photocatalytic degradation of ...

Source Fuller (1966) [FSG66], Tucker & Nelken (1982) [TN82]

Parameters

• mol_volume (float) – molar volume [cm^3/mol]

• molar_weight (float) – molar weight [g/mol]

• temperature (float, int, optional) – temperature [K]

• pressure (float, optional) – pressure [Pa]

Returns 𝐷 - diffusion constant [cm^2/s]

Return type float

classmethod get_diffusion_constant_water(mol_volume, temperature=293.15,model='s')

Diffusion constant in water depending on molar volume. Based on models from either Stokes, Wilke-Chang, or Hayduk-Minhas.

Source Wilke & Chang (1955) [WC55], Hayduk & Minhas (1982) [HM82]

Parameters

• mol_volume (float) – molar volume [cm^3/mol]

• temperature (float, int, optional) – temperature [K]

• model (str, optional) – model abbreviation | “s” Stokes (default), “wc” Wilke-Chang, “hm” Hayduk-Minhas

Returns 𝐷 - diffusion constant [m^2/s]

Return type float

classmethod get_molecular_weight(composition)Calculate the weight of a molecule from its chemical formula.

Parameters composition (dict) – key - atomic symbol | value - atom count

Returns molecular weight [u]

Return type float

k_boltzmann = 1.3806488e-23𝑘𝐵 - Boltzmann constant [Nm/K]

light_speed = 2997927458𝑐 - light speed [m/s]

planck = 6.62606957e-34ℎ - Planck constant [Ws^2]

static scale_ten(value)Decimal scaling

Parameters value (int, float) – input value

Returns scale factor, prefix

Return type tuple

static scale_time(value)Temporal scaling

28

132

Page 133: Modeling and simulation of photocatalytic degradation of ...

Parameters value (int, float) – input value

Returns scale factor, prefix

Return type tuple

standard_atomic_weight = {'Ac': 225.027, 'Ag': 107.87, ...}Mapping element symbol to standard corresponding atomic weight.

Source “Atomic weights of the elements 2013” (IUPAC) [MCB+16]

Returns standard atomic weight [u]

Return type dict

classmethod unit_convert(value, unit)Convert into pdom base units

Note Base units are s, 1/s, m, m/s, W/m^2, K, g/m^3, m^2/g, m^3, cm^3/mol, m^2/molecule, kJ/mol

Parameters

• value (float, int) –

• unit (str) –

Returns converted value

Return type float

unit_convert_concentration(value, unit, cfg=None)Concentration to number of molecules

Note A simulation config can be passed, to use alternative parameters, for example to con-vert TOC.

Parameters

• value (float, int) – value

• unit (str) – unit

• cfg (dict, optional) – simulation config

Returns 𝑁 - number of molecules [1]

Return type float

static update_multi_k(config)Update the config for multi species models

This needs to be called after a basic setting is changed in the config, to update parameters dependingon species size.

Parameters config (dict) – simulation config

Returns simulation config

Return type dict

van_der_waals_radii = {'Ac': 2.0, 'Ag': 1.72, 'Al': 2.0, ...}Mapping element symbol to corresponding Van der Waals radius.

Note Radii that are not available in either of the sources have RvdW = 2.00.

29

133

Page 134: Modeling and simulation of photocatalytic degradation of ...

Source “van der Waals Volumes and Radii” by Bondi (1964) [Bon64], the value for H istaken from Rowland & Taylor (1996) [RT96]

Returns Van der Waals radii [Ang]

Return type dict

viscosity_water = <scipy.interpolate.interpolate.interp1d object>A smooth interpolation from 274 to 363 Kelvin. Viscosity is retuned in mPa*s | centipoise |(mN/m^2)*s.

Source Wikibooks - Stoffdaten Wasser

5.3 pdom.Molecule

class pdom.Molecule(identifier, properties, save=True)This class should not be initialised directly. Use one of the following class methods in-stead: from_chem_id(), from_folder(), from_inchi(), from_inchi_key(),from_iupac_name(), from_name()

Parameters

• identifier (dict) –

– name common name

– iupac_name IUPAC name (can be the same as common name)

– chem_id compound ID from PubChem

– inchi IUPAC International Chemical Identifier (human readable)

– inchi_key IUPAC International Chemical Identifier (hash)

• properties (dict) –

– chem_formula chemical formula e.g. ‘C10H22’

– chem_formdic chemical formula as dict

– mol_weight molecular weight in g/mol

– mol_volume molecular volume in cm^3/mol

– mol_surface largest projected surface area m^2/molecule

– excess_bonds number of bonds in excess of a simple carbon chain

– structure_3d atom position list

• save (bool) – save data tu the user cache

classmethod from_chem_id(chem_id, name=None, inchi_key=None)Create Molecule instance identified by a chemID from PubChem data.

Note if molecule is cached in Molecule.db_folder load from there

Parameters

• chem_id (int) – compound ID from PubChem

• name (str, optional) – overwrite compound name

30

134

Page 135: Modeling and simulation of photocatalytic degradation of ...

• inchi_key (str, optional) – IUPAC International Chemical Identifier (tocheck cache quickly)

Returns Molecule instance

classmethod from_folder(folder)create Molecule instance from a folder created by save

Parameters folder (str, folder) – molecule folder

Returns Molecule instance

classmethod from_name(name)Create Molecule instance identified by a name if unique

Note queries chem_id and calls from_chem_id()

Parameters name (str) – unique compound name

Returns Molecule instance

identifierdict of identifiers as described in Molecule

propertiesdict of properties as described in Molecule

save(folder=None, name=None)saves the molecule information to disk, can be loaded with from_folder()

Note If called without parameters the molecule is saved in an appropriate cache folderMolecule.db_folder.

Parameters

• folder (str, Path) – parent folder

• name (str) – molecule folder name (default INCHI key)

structure_3d(rotated=True)Returned 3d structure of the molecule

Parameters rotated (bool) – if True the structure is rotated to cover the maximumsurface in the xy space

Returns 3D structure (symbol, x, y, z)

Return type list

5.4 pdom.export

Helper functions to export simulation results.

pdom.export.fit_dark(cfg, result, result_raw)Export function for fit: Adsorption - Desorption experiment in the dark.

Parameters

• cfg (dict) – simulation config

• result (tuple) – fit_t, fit_y, k_ads, k_des

• result_raw – raw fit function result

31

135

Page 136: Modeling and simulation of photocatalytic degradation of ...

Result

• plot with fit and initial data points fit_dark.pdf

• space separated data file containing the fit fit_dark-values_calculated.txt

• space separated data file containing initial data pointsfit_dark-values_experiment.txt

• .json with the fitted parameters fit_dark.json

• .txt files with corresponding units (LaTeX formatted)

pdom.export.fit_single(cfg, result, result_raw)Export function for fit: single species degradation model.

Parameters

• cfg (dict) – simulation config

• result (tuple) – fit_t, fit_y, k_ads, k_des, k_reac

• result_raw – raw fit function result

Result

• plot with fit and initial data points fit_single.pdf

• space separated data file containing the fit fit_single-values_calculated.txt

• space separated data file containing initial data pointsfit_single-values_experiment.txt

• .json with the fitted parameters fit_single.json

• .txt files with corresponding units (LaTeX formatted)

pdom.export.fit_toc(cfg, sd_error, t, fit)Export function for fit: TOC.

Parameters

• cfg (dict) – simulation config

• sd_error (float) – standard derivation error

• t (ndarray) – fit time series

• fit (ndarray) – fit toc series

Result

• plot with fit and initial data points | fit_toc.pdf

• space separated data file containing the fit | fit_toc-values_calculated.txt

• space separated data file containing initial data points |fit_toc-values_experiment.txt

• .json with the fitted parameters | fit_toc.json

• .txt files with corresponding units (LaTeX formatted)

pdom.export.multi_species(cfg, t, N_surf, N_vol)Export function for incremental and fragmentation multi species model simulations.

32

136

Page 137: Modeling and simulation of photocatalytic degradation of ...

Parameters

• cfg (dict) – simulation config

• t (ndarray) – fit time series

• N_surf (ndarray) – number of molecules on the surface

• N_vol (ndarray) – number of molecules in solution

Result

• plot with concentration development in solution and TOC c_volume-toc.pdf

• plot of the segments in solution and on the surface volume_segments.pdf,solution_segments.pdf

• space separated data file containing the concentrations in solutionmulti_species-values_volume.txt

• space separated data file containing the concentrations on the surfacemulti_species-values_surface.txt

• .txt files with corresponding units (LaTeX formatted)

pdom.export.multi_species_bonds(cfg, t, N_surf_detail, N_vol_detail)Export function for incremental and fragmentation multi species model simulations.

Parameters

• cfg (dict) – simulation config

• t (ndarray) – fit time series

• N_surf_detail (ndarray) – number of molecules on the surface

• N_vol_detail (ndarray) – number of molecules in solution

Result

• same files as multi_species()

• plot of the average excess bond count excess_bonds.pdf

• space separated data file containing the average number of excess bondsmulti_species-excess_bonds

pdom.export.single_species(cfg, t, N_surf, N_vol)Export function single species model simulations.

Parameters

• cfg (dict) – simulation config

• t (ndarray) – fit time series

• N_surf (ndarray) – number of molecules on the surface

• N_vol (ndarray) – number of molecules in solution

Result

• plot with concentration development in solution and on the surfacesingle_species.pdf

• space separated data file containing the concentrations in solutionsingle_species-values_volume.txt

33

137

Page 138: Modeling and simulation of photocatalytic degradation of ...

• space separated data file containing the concentrations on the surfacesingle_species-values_surface.txt

• .txt files with corresponding units (LaTeX formatted)

References

[Bon64] A. Bondi. van der Waals Volumes and Radii. The Journal of Physical Chemistry, 68(3):441–451,1964. doi:10.1021/j100785a001.

[EBT+15] Hagen Eckert, Manfred Bobeth, Sara Teixeira, Klaus Kühn, and Gianaurelio Cuniberti. Modelingof photocatalytic degradation of organic components in water by nanoparticle suspension. ChemicalEngineering Journal, 261:67–75, 2015. doi:10.1016/j.cej.2014.05.147.

[FSG66] Edward N. Fuller, Paul D. Schettler, and J. Calvin Giddings. New method for prediction of bi-nary gas-phase diffusion coefficients. Industrial & Engineering Chemistry, 58(5):18–27, 1966.doi:10.1021/ie50677a007.

[HM82] W. Hayduk and B. S. Minhas. Correlations for prediction of molecular diffusivities in liquids. TheCanadian Journal of Chemical Engineering, 60(2):295–299, 1982. doi:10.1002/cjce.5450600213.

[Hou01] Ammar Houas. Photocatalytic degradation pathway of methylene blue in water. Applied Catalysis B:Environmental, 31(2):145–157, 2001. doi:10.1016/S0926-3373(00)00276-9.

[MCB+16] Juris Meija, Tyler B. Coplen, Michael Berglund, Willi A. Brand, Paul De Bièvre, Manfred Grön-ing, Norman E. Holden, Johanna Irrgeher, Robert D. Loss, Thomas Walczyk, and Thomas Pro-haska. Atomic weights of the elements 2013 (IUPAC Technical Report). Pure and Applied Chemistry,88(3):265–291, 2016. doi:10.1515/pac-2015-0305.

[RT96] R. Scott Rowland and Robin Taylor. Intermolecular Nonbonded Contact Distances in Organic Crys-tal Structures: Comparison with Distances Expected from van der Waals Radii. Journal of PhysicalChemistry, 100(18):7384–7391, 1996. doi:10.1021/jp953141+.

[TN82] William A. Tucker and Leslie H. Nelken. Diffusion Coefficients in Air and Water. In Warren J.Lyman, William F. Reehl, and David H. Rosenblatt, editors, Handbook of chemical property es-timation methods : environmental behavior of organic compounds. McGraw-Hill, 1982. URL:https://archive.org/details/handbookofchemic0000lyma.

[WC55] C. R. Wilke and Pin Chang. Correlation of diffusion coefficients in dilute solutions. AIChE Journal,1(2):264–270, 1955. doi:10.1002/aic.690010222.

34

138

Page 139: Modeling and simulation of photocatalytic degradation of ...

List of figures

2.1. Relationship between the model and the real and virtual systems. . . . . 21

2.2. Energy diagram for a typical n-type semiconductor. . . . . . . . . . . . . 24

2.3. Band structure of the anatase and rutile phase of TiO2. . . . . . . . . . . 25

2.4. Energy diagram for a typical n-type semiconductor in contact with an

electrolyte in dark and under illumination. . . . . . . . . . . . . . . . . . . 28

2.5. Bandgap energies and absorption edges for selected metal oxides. . . . 32

2.6. Bandgap energies and absorption edges for selected chalcogenides and

other binary compounds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

2.7. Bandgap energies and absorption edges for selected ternary and qua-

ternary semiconductors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

2.8. Crystal structure of rutile and anatase from different perspectives. . . . 34

2.9. Crystal structure of zinc oxide - wurtzite from different perspectives. . . 37

2.10.Long-term average of direct horizontal solar irradiation for central Eu-

rope and south-east Asia. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3.1. Experimental setup for the photocatalytic degradation experiments. . . 52

3.2. Wastewater treatment plant in Kaditz operated by Stadtentwässerung

Dresden GmbH. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

3.3. Illustration of processes occurring during mineralization of organic mol-

ecules at the surface of photocatalytic nanoparticles in solution. . . . . . 57

3.4. Steps to calculate themaximal projected accessible surface area of cipro-

floxacin. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

3.5. Illustration of the reaction steps in the two basicmultiple organic species

models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

139

Page 140: Modeling and simulation of photocatalytic degradation of ...

3.6. Structure of ciprofloxacin. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

3.7. Bond model reaction schematics for ciprofloxacin. . . . . . . . . . . . . . 70

4.1. SEM images of colloidal titanium dioxide (TiO2) and zinc oxide (ZnO). . . 77

4.2. Light output of the usedmercury fluorescent tubes normalized at 365nm

compared to the diffuse reflectance spectra of TiO2 and ZnO. . . . . . . 78

4.3. Simulation results for the adsoprtion/desorption dynamic of methylene

blue on TiO2 based on data reported in [159] . . . . . . . . . . . . . . . . 80

4.4. Measured concentration evolutions during the degradation of ciproflox-

acin and methylene blue under UV-irradiation with 1.0 g L−1 catalysts

TiO2 and ZnO. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

4.5. Measured concentration evolutions during degradation of ciprofloxacin

with TiO2. Compared to Fig. 4.4 a), a reduced TiO2 particle concentration

of 0.5 g L−1, and a lower UV-irradiation of 5.5Wm−2 was used. . . . . . . 83

4.6. Calculated concentration evolutions of intermediates based on the in-

cremental oxidation or fragmentation model with weak or strong de-

pendence of the desorption rate constants on the molecule size n. . . . 84

4.7. Change in the overall excess bonds and average number of excess bonds

for intermediates of size n. . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

4.8. Comparison between the TOC evolution for the incremental, the frag-

mentation, and the excess bonds oxidation model for weak and strong

size dependence of the desorption rate. . . . . . . . . . . . . . . . . . . . 87

4.9. Simulation results for the degradation of methylene blue by TiO2 based

on data reported in [159]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

4.10.Measured concentration during the degradation of carbamazepine un-

der UV irradiation with 1.0 g L−1 catalysts TiO2 and ZnO. . . . . . . . . . . 91

4.11.Average degradation ratio over time of the 14 selected pharmaceuticals

measured by HPLC-MS/MS-method. . . . . . . . . . . . . . . . . . . . . . 92

140

Page 141: Modeling and simulation of photocatalytic degradation of ...

B.1. Concentration evolutionsmeasured byHPLC-MS/MS-methodduring the

degradation of different pharmaceuticals with low initial concentration

under UV-irradiation with 1 g L−1 catalysts. . . . . . . . . . . . . . . . . . . 103

B.2. Concentration evolutions continued from Fig. B.1. . . . . . . . . . . . . . 104

List of tables

3.1. Standard parameters measured in the effluent of the wastewater treat-

ment plant in Kaditz. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

3.2. Selected pharmaceuticals from different drug classes to be monitored

in the degradation experiment. . . . . . . . . . . . . . . . . . . . . . . . . 55

4.1. Adsorption and desorption rate constants derived from experiments

in the dark for the compounds ciprofloxacin and methylene blue using

photocatalysts TiO2 and ZnO, respectively. . . . . . . . . . . . . . . . . . . 79

4.2. Adsorption anddesorption rate constants derived from the experiments

with carbamazepine using photocatalytic TiO2 and ZnO nanoparticles. . 90

4.3. Apparent rate constants average and adsorption rate constants aver-

age for the 14 studied pharmaceuticals (Tab. 3.2) with the photocatalysts

TiO2 and ZnO. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

141

Page 142: Modeling and simulation of photocatalytic degradation of ...

Bibliography

[1] P. J. Mohr, D. B. Newell, and B. N. Taylor.

“CODATA recommended values of the fundamental physical constants: 2014”.

In: Reviews of Modern Physics 88.3, 035009 (2016).doi: 10.1103/RevModPhys.88.035009.

[2] C. D. Metcalfe, B. G. Koenig, D. T. Bennie, M. Servos, T. Ternes, and R. Hirsch.

“Occurrence of neutral and acidic drugs in the effluents of Canadian sewage

treatment plants.”

In: Environmental Toxicology and Chemistry 22.12 (2003), pp. 2872–2880.doi: 10.1897/02-469.

[3] M. Isidori, M. Bellotta, M. Cangiano, and A. Parrella.

“Estrogenic activity of pharmaceuticals in the aquatic environment.”

In: Environment international 35.5 (2009), pp. 826–829.doi: 10.1016/j.envint.2008.11.006.

[4] M. Klavarioti, D. Mantzavinos, and D. Kassinos. “Removal of residual

pharmaceuticals from aqueous systems by advanced oxidation processes.”

In: Environment international 35.2 (2009), pp. 402–417.doi: 10.1016/j.envint.2008.07.009.

[5] N. Xekoukoulotakis, N. Xinidis, M. Chroni, et al.

“UV-A/TiO2 photocatalytic decomposition of erythromycin in water: Factors

affecting mineralization and antibiotic activity”.

In: Catalysis Today 151.1-2 (2010), pp. 29–33.doi: 10.1016/j.cattod.2010.01.040.

[6] V. Cunningham, S. Binks, and M. Olson. “Human health risk assessment from

the presence of human pharmaceuticals in the aquatic environment.”

In: Regulatory toxicology and pharmacology 53.1 (2009), pp. 39–45.doi: 10.1016/j.yrtph.2008.10.006.

[7] J. Pereira, V. Vilar, M. Borges, O. González, S. Esplugas, and R. Boaventura.

“Photocatalytic degradation of oxytetracycline using TiO2 under natural and

simulated solar radiation”. In: Solar Energy 85.11 (2011), pp. 2732–2740.doi: 10.1016/j.solener.2011.08.012.

142

Page 143: Modeling and simulation of photocatalytic degradation of ...

[8] O. A. H. Jones, N. Voulvoulis, and J. N. Lester. “Aquatic environmental

assessment of the top 25 English prescription pharmaceuticals.”

In: Water research 36.20 (2002), pp. 5013–5022.doi: 10.1016/s0043-1354(02)00227-0.

[9] W. C. Li. “Occurrence, sources, and fate of pharmaceuticals in aquatic

environment and soil.”

In: Environmental pollution (Barking, Essex : 1987) 187 (2014), pp. 193–201.doi: 10.1016/j.envpol.2014.01.015.

[10] W. Mrozik and J. Stefańska.

“Adsorption and biodegradation of antidiabetic pharmaceuticals in soils.”

In: Chemosphere 95 (2014), pp. 281–288.doi: 10.1016/j.chemosphere.2013.09.012.

[11] T. Ternes, M. Meisenheimer, D. McDowell, et al.

“Removal of pharmaceuticals during drinking water treatment.”

In: Environmental science & technology 36.17 (2002), pp. 3855–3863.doi: 10.1021/es015757k.

[12] M. S. Fram and K. Belitz.

“Occurrence and concentrations of pharmaceutical compounds in

groundwater used for public drinking-water supply in California.”

In: The Science of the total environment 409.18 (2011), pp. 3409–3417.doi: 10.1016/j.scitotenv.2011.05.053.

[13] T. E. Doll and F. H. Frimmel.

“Fate of pharmaceuticals – photodegradation by simulated solar UV-light.”

In: Chemosphere 52.10 (2003), pp. 1757–1769.doi: 10.1016/S0045-6535(03)00446-6.

[14] N. Xekoukoulotakis, C. Drosou, C. Brebou, et al.

“Kinetics of UV-A/TiO2 photocatalytic degradation and mineralization of the

antibiotic sulfamethoxazole in aqueous matrices”.

In: Catalysis Today 161.1 (2011), pp. 163–168.doi: 10.1016/j.cattod.2010.09.027.

[15] Y. Lu, H. Patricia, A. Dele, and E. Tessa. “Medicine Expenditures”.

In: The World Medicines Situation 2011. 3rd ed.World Health Organization, 2011, pp. 1–34.

143

Page 144: Modeling and simulation of photocatalytic degradation of ...

[16] C. G. Daughton and B. W. Brooks.

“Active Pharmaceutical Ingredients and Aquatic Organisms”.

In: Environmental Contaminants in Biota: Interpreting Tissue Concentrations.Ed. by B. W. Nelson. 2nd ed. CRC Press, 2011, pp. 255–280.

doi: 10.1201/b10598.

[17] B. Halling-Sørensen, S. Nors Nielsen, P. Lanzky, F. Ingerslev, H. Holten Lützhøft,

and S. Jørgensen. “Occurrence, fate and effects of pharmaceutical substances

in the environment- A review”. In: Chemosphere 36.2 (1998), pp. 357–393.doi: 10.1016/S0045-6535(97)00354-8.

[18] M. D. Hernando, M. Mezcua, A. R. Fernández-Alba, and D. Barceló.

“Environmental risk assessment of pharmaceutical residues in wastewater

effluents, surface waters and sediments.” In: Talanta 69.2 (2006), pp. 334–342.doi: 10.1016/j.talanta.2005.09.037.

[19] I. Michael, L. Rizzo, C. S. McArdell, et al. “Urban wastewater treatment plants as

hotspots for the release of antibiotics in the environment: a review.”

In: Water research 47.3 (2013), pp. 957–995. doi: 10.1016/j.watres.2012.11.027.[20] WTO. World Trade Statistical Review 2018.

Geneva, Switzerland: World Trade Organization, 2018.

[21] K. T. Chung, G. E. Fulk, and A. W. Andrews.

“Mutagenicity testing of some commonly used dyes.”

In: Applied and environmental microbiology 42.4 (1981), pp. 641–648.[22] R. O. Alves de Lima, A. P. Bazo, D. M. F. Salvadori, C. M. Rech, D. de Palma

Oliveira, and G. de Aragão Umbuzeiro.

“Mutagenic and carcinogenic potential of a textile azo dye processing plant

effluent that impacts a drinking water source.”

In: Mutation research 626.1-2 (2007), pp. 53–60.doi: 10.1016/j.mrgentox.2006.08.002.

[23] Council of European Union. “Regulation (EC) No 1907/2006 of the European

Parliament and of the Council of 18 December 2006 concerning the

Registration, Evaluation, Authorisation and Restriction of Chemicals (REACH),

establishing a European Chemicals Agency, amending Directive 1999/4”.

In: Official Journal of the European Union 49, OJ:L396 (2006), pp. 1–850.[24] H. Zollinger. Color chemistry: syntheses, properties, and applications of organic

dyes and pigments. 3rd ed. Weinheim: Wiley-VCH, 2003.

144

Page 145: Modeling and simulation of photocatalytic degradation of ...

[25] R. Anliker. “Color chemistry and the environment”.

In: Ecotoxicology and Environmental Safety 1.2 (1977), pp. 211–237.doi: 10.1016/0147-6513(77)90037-9.

[26] R. Andreozzi, V. Caprio, A. Insola, and R. Marotta.

“Advanced oxidation processes (AOP) for water purification and recovery”.

In: Catalysis Today 53 (1999), pp. 51–59. doi: 10.1016/S0920-5861(99)00102-9.[27] A. Ray and A. Beenackers.

“Development of a new photocatalytic reactor for water purification”.

In: Catalysis Today 40 (1998), pp. 73–83. doi: 10.1016/S0920-5861(97)00123-5.[28] L. Yang, L. Yu, and M. Ray.

“Degradation of paracetamol in aqueous solutions by TiO2 photocatalysis.”

In: Water research 42.13 (2008), pp. 3480–3488.doi: 10.1016/j.watres.2008.04.023.

[29] E. Hapeshi, A. Achilleos, M. Vasquez, et al.

“Drugs degrading photocatalytically: Kinetics and mechanisms of ofloxacin and

atenolol removal on titania suspensions.”

In: Water research 44.6 (2010), pp. 1737–1746.doi: 10.1016/j.watres.2009.11.044.

[30] A. Fujishima, T. N. Rao, and D. A. Tryk. “Titanium dioxide photocatalysis”.

In: Journal of Photochemistry and Photobiology C: Photochemistry Reviews 1.1(2000), pp. 1–21. doi: 10.1016/S1389-5567(00)00002-2.

[31] F. Thevenet, O. Guaïtella, J. Herrmann, a. Rousseau, and C. Guillard.

“Photocatalytic degradation of acetylene over various titanium dioxide-based

photocatalysts”. In: Applied Catalysis B: Environmental 61.1-2 (2005), pp. 58–68.doi: 10.1016/j.apcatb.2005.03.015.

[32] M. Faramarzpour, M. Vossoughi, and M. Borghei. “Photocatalytic degradation

of furfural by titania nanoparticles in a floating-bed photoreactor”.

In: Chemical Engineering Journal 146.1 (2009), pp. 79–85.doi: 10.1016/j.cej.2008.05.033.

[33] A. Kathiravan, M. Asha Jhonsi, and R. Renganathan.

“Photoinduced interaction of colloidal TiO2 nanoparticles with lysozyme:

Evidences from spectroscopic studies”.

In: Journal of Luminescence 131.9 (2011), pp. 1975–1981.doi: 10.1016/j.jlumin.2011.04.004.

145

Page 146: Modeling and simulation of photocatalytic degradation of ...

[34] D. S. Bhatkhande, V. G. Pangarkar, and A. A. Beenackers.

“Photocatalytic degradation for environmental applications - a review”.

In: Journal of Chemical Technology & Biotechnology 77.1 (2002), pp. 102–116.doi: 10.1002/jctb.532.

[35] I. K. Konstantinou and T. Albanis. “TiO2-assisted photocatalytic degradation of

azo dyes in aqueous solution: kinetic and mechanistic investigations”.

In: Applied Catalysis B: Environmental 49.1 (2004), pp. 1–14.doi: 10.1016/j.apcatb.2003.11.010.

[36] M. Faramarzpour, M. Vossoughi, and M. Borghei. “Photocatalytic degradation

of furfural by titania nanoparticles in a floating-bed photoreactor”.

In: Chemical Engineering Journal 146.1 (2009), pp. 79–85.doi: 10.1016/j.cej.2008.05.033.

[37] L. Hu, P. M. Flanders, P. L. Miller, and T. J. Strathmann. “Oxidation of

sulfamethoxazole and related antimicrobial agents by TiO2 photocatalysis.”

In: Water research 41.12 (2007), pp. 2612–2626.doi: 10.1016/j.watres.2007.02.026.

[38] M. Lazar, S. Varghese, and S. Nair.

“Photocatalytic Water Treatment by Titanium Dioxide: Recent Updates”.

In: Catalysts 2.4 (2012), pp. 572–601. doi: 10.3390/catal2040572.[39] A. Fujishima, X. Zhang, and Donald A. Trykc.

“TiO2 photocatalysis and related surface phenomena”.

In: Surface Science Reports 63.12 (2008), pp. 515–582.doi: 10.1016/j.surfrep.2008.10.001.

[40] L. Pan, X. Liu, Z. Sun, and C. Q. Sun. “Nanophotocatalysts via

microwave-assisted solution-phase synthesis for efficient photocatalysis”.

In: Journal of Materials Chemistry A 1.29 (2013), p. 8299.doi: 10.1039/c3ta10981j.

[41] L. K. Adams, D. Y. Lyon, and P. J. Alvarez.

“Comparative eco-toxicity of nanoscale TiO2, SiO2, and ZnO water suspensions”.

In: Water Research 40.19 (2006), pp. 3527–3532.doi: 10.1016/j.watres.2006.08.004.

[42] A. E. Cassano, C. A. Martin, R. J. Brandi, and O. M. Alfano.

“Photoreactor Analysis and Design: Fundamentals and Applications”.

In: Industrial & Engineering Chemistry Research 34.7 (1995), pp. 2155–2201.doi: 10.1021/ie00046a001.

146

Page 147: Modeling and simulation of photocatalytic degradation of ...

[43] M. Motegh, J. Cen, P. W. Appel, J. R. van Ommen, and M. T. Kreutzer.

“Photocatalytic-reactor efficiencies and simplified expressions to assess their

relevance in kinetic experiments”.

In: Chemical Engineering Journal 207-208 (2012), pp. 607–615.doi: 10.1016/j.cej.2012.07.023.

[44] V. Vaiano, O. Sacco, M. Stoller, A. Chianese, P. Ciambelli, and D. Sannino.

“Influence of the Photoreactor Configuration and of Different Light Sources in

the Photocatalytic Treatment of Highly Polluted Wastewater”.

In: International Journal of Chemical Reactor Engineering 12.1 (2014), pp. 1–13.doi: 10.1515/ijcre-2013-0090.

[45] M. N. Chong, B. Jin, C. W. K. Chow, and C. Saint.

“Recent developments in photocatalytic water treatment technology: a review.”

In: Water research 44.10 (2010), pp. 2997–3027.doi: 10.1016/j.watres.2010.02.039.

[46] C. Minero. “Kinetic analysis of photoinduced reactions at the water

semiconductor interface”. In: Catalysis Today 54.2-3 (1999), pp. 205–216.doi: 10.1016/S0920-5861(99)00183-2.

[47] R. Andreozzi, V. Caprio, A. Insola, G. Longo, and V. Tufano.

“Photocatalytic oxidation of 4-nitrophenol in aqueous TiO2 slurries: an

experimental validation of literature kinetic models”.

In: Journal of Chemical Technology & Biotechnology 75.2 (2000), pp. 131–136.doi: 10.1002/(SICI)1097-4660(200002)75:2<131::AID-JCTB191>3.0.CO;2-F.

[48] D. Sannino, V. Vaiano, O. Sacco, and P. Ciambelli. “Mathematical modelling of

photocatalytic degradation of methylene blue under visible light irradiation”.

In: Journal of Environmental Chemical Engineering 1.1-2 (2013), pp. 56–60.doi: 10.1016/j.jece.2013.03.003.

[49] C.-H. Wu and J.-M. Chern.

“Kinetics of Photocatalytic Decomposition of Methylene Blue”.

In: Industrial & Engineering Chemistry Research 45.19 (2006), pp. 6450–6457.doi: 10.1021/ie0602759.

[50] D. E. Rumelhart, G. E. Hinton, and R. J. Williams.

“Learning representations by back-propagating errors”.

In: Nature 323.6088 (1986), pp. 533–536. doi: 10.1038/323533a0.

147

Page 148: Modeling and simulation of photocatalytic degradation of ...

[51] B. J. Alder and T. E. Wainwright.

“Studies in Molecular Dynamics. I. General Method”.

In: The Journal of Chemical Physics 31.2 (1959), pp. 459–466.doi: 10.1063/1.1730376.

[52] A. Rahman. “Correlations in the Motion of Atoms in Liquid Argon”.

In: Physical Review 136.2A (1964), A405–A411. doi: 10.1103/PhysRev.136.A405.

[53] M. Levitt and A. Warshel. “Computer simulation of protein folding”.

In: Nature 253.5494 (1975), pp. 694–698. doi: 10.1038/253694a0.[54] M. Levitt and R. Sharon. “Accurate simulation of protein dynamics in solution.”

In: Proceedings of the National Academy of Sciences 85.20 (1988), pp. 7557–7561.doi: 10.1073/pnas.85.20.7557.

[55] Y. Duan and A. Kollman. “Pathways to a Protein Folding Intermediate Observed

in a 1-Microsecond Simulation in Aqueous Solution”.

In: Science 282.5389 (1998), pp. 740–744. doi: 10.1126/science.282.5389.740.[56] P. L. Freddolino, A. S. Arkhipov, S. B. Larson, A. McPherson, and K. Schulten.

“Molecular Dynamics Simulations of the Complete Satellite Tobacco Mosaic

Virus”. In: Structure 14.3 (2006), pp. 437–449. doi: 10.1016/j.str.2005.11.014.[57] P. Schroll. “Early pioneers of organic photochemistry”.

In: Chemical Photocatalysis. Ed. by B. König. Berlin, Boston: DE GRUYTER, 2013,pp. 3–18. doi: 10.1515/9783110269246.3.

[58] L. Bruner and J. Kozak. “Zur Kenntnis der Photokatalyse. I. Die Lichtreaktion in

Gemischen: Uransalz + Oxalsäure”.

In: Zeitschrift für Elektrochemie 17.9 (1911), pp. 354–360.doi: 10.1002/bbpc.19110170905.

[59] E. C. C. Baly, I. M. Heilbron, and W. F. Barker. “Photocatalysis. Part I. The

synthesis of formaldehyde and carbohydrates from carbon dioxide and water”.

In: J. Chem. Soc., Trans. 119 (1921), pp. 1025–1035.doi: 10.1039/CT9211901025.

[60] E. C. C. Baly, I. M. Heilbron, and D. P. Hudson. “Photocatalysis. Part II. The

photosynthesis of nitrogen compounds from nitrates and carbon dioxide”.

In: J. Chem. Soc., Trans. 121 (1922), pp. 1078–1088.doi: 10.1039/CT9222101078.

148

Page 149: Modeling and simulation of photocatalytic degradation of ...

[61] S. E. Braslavsky, A. M. Braun, A. E. Cassano, et al. “Glossary of terms used in

photocatalysis and radiation catalysis (IUPAC Recommendations 2011)”.

In: Pure and Applied Chemistry 83.4 (2011), pp. 931–1014.doi: 10.1351/PAC-REC-09-09-36.

[62] A. Fujishima and K. Honda.

“Electrochemical Photolysis of Water at a Semiconductor Electrode”.

In: Nature 238.5358 (1972), pp. 37–38. doi: 10.1038/238037a0.[63] G. Ciamician. “The Photochemistry of the Future”.

In: Science 36.926 (1912), pp. 385–394. doi: 10.1126/science.36.926.385.[64] A. Pfitzner, S. Dankesreiter, A. Eisenhofer, and M. Cherevatskaya.

“12 Heterogeneous semiconductor photocatalysis”. In: Chemical Photocatalysis.Ed. by B. König. Berlin, Boston: DE GRUYTER, 2013.

Chap. Heterogene, pp. 211–246. doi: 10.1515/9783110269246.211.

[65] R. H. Taylor, F. Rose, C. Toher, et al.

“A RESTful API for exchanging materials data in the AFLOWLIB.org consortium”.

In: Computational Materials Science 93 (2014), pp. 178–192.doi: 10.1016/j.commatsci.2014.05.014.

[66] Y. Tezuka, S. Shin, T. Ishii, T. Ejima, S. Suzuki, and S. Sato. “Photoemission and

Bremsstrahlung Isochromat Spectroscopy Studies of TiO2 (Rutile) and SrTiO3”.

In: Journal of the Physical Society of Japan 63.1 (1994), pp. 347–357.doi: 10.1143/JPSJ.63.347.

[67] L. Kavan, M. Grätzel, S. E. Gilbert, C. Klemenz, and H. J. Scheel. “Electrochemical

and Photoelectrochemical Investigation of Single-Crystal Anatase”.

In: Journal of the American Chemical Society 118.28 (1996), pp. 6716–6723.doi: 10.1021/ja954172l.

[68] H. Gerischer and W. Ekardt.

“Fermi levels in electrolytes and the absolute scale of redox potentials”.

In: Applied Physics Letters 43.4 (1983), pp. 393–395. doi: 10.1063/1.94356.[69] S. Trasatti. “The absolute electrode potential: an explanatory note

(Recommendations 1986)”.

In: Pure and Applied Chemistry 58.7 (1986), pp. 955–966.doi: 10.1351/pac198658070955.

149

Page 150: Modeling and simulation of photocatalytic degradation of ...

[70] L. M. Peter. “Photoelectrochemistry: From Basic Principles to Photocatalysis”.

In: Photocatalysis: Fundamentals and Perspectives.Ed. by J. Schneider, D. Bahnemann, J. Ye, G. Li Puma, and D. D. Dionysiou.

Energy and Environment Series. Cambridge: Royal Society of Chemistry, 2016,

pp. 1–28. doi: 10.1039/9781782622338-00001.

[71] E. A. Davis and N. F. Mott.

“Conduction in non-crystalline systems V. Conductivity, optical absorption and

photoconductivity in amorphous semiconductors”.

In: Philosophical Magazine 22.179 (1970), pp. 903–922.doi: 10.1080/14786437008221061.

[72] B. D. Viezbicke, S. Patel, B. E. Davis, and D. P. Birnie.

“Evaluation of the Tauc method for optical absorption edge determination:

ZnO thin films as a model system”.

In: physica status solidi (b) 252.8 (2015), pp. 1700–1710.doi: 10.1002/pssb.201552007.

[73] L. M. Peter and D. Vanmaekelbergh.

“Time and Frequency Resolved Studies of Photoelectrochemical Kinetics”.

In: Advances in Electrochemical Science and Engineering, Volume 6.Ed. by R. C. Alkire and D. M. Kolb. Vol. 2. 2008, pp. 77–163.

doi: 10.1002/9783527616800.ch2.

[74] L. Peter. “Chapter 8 Photocurrent Spectroscopy”.

In: Comprehensive Chemical Kinetics. Vol. 29. 1989, pp. 353–383.doi: 10.1016/S0069-8040(08)70323-6.

[75] Krishnan Rajeshwar. “Fundamentals of Semiconductor Electrochemistry and

Photoelectrochemistry”. In: Encyclopedia of Electrochemistry.Weinheim, Germany: Wiley-VCH Verlag GmbH & Co. KGaA, 2007, pp. 1–53.

doi: 10.1002/9783527610426.bard060001.

[76] W. W. Gärtner. “Depletion-Layer Photoeffects in Semiconductors”.

In: Physical Review 116.1 (1959), pp. 84–87. doi: 10.1103/PhysRev.116.84.

[77] A. Hernández-Ramírez and I. Medina-Ramírez. Photocatalytic Semiconductors.2015, pp. 1–289. doi: 10.1007/978-3-319-10999-2.

[78] S. G. Botta, J. A. Navío, M. C. Hidalgo, G. M. Restrepo, and M. I. Litter.

“Photocatalytic properties of ZrO2 and Fe/ZrO2 semiconductors prepared by a

sol–gel technique”. In: Journal of Photochemistry and Photobiology A: Chemistry129.1-2 (1999), pp. 89–99. doi: 10.1016/S1010-6030(99)00150-1.

150

Page 151: Modeling and simulation of photocatalytic degradation of ...

[79] J. Navío, G. Colón, M. Macías, P. Sánchez-Soto, V. Augugliaro, and L. Palmisano.

“ZrO2–SiO2 mixed oxides: surface aspects, photophysical properties and

photoreactivity for 4-nitrophenol oxidation in aqueous phase”.

In: Journal of Molecular Catalysis A: Chemical 109.3 (1996), pp. 239–248.doi: 10.1016/1381-1169(96)00009-X.

[80] G. H. He, C. J. Liang, Y. D. Ou, D. N. Liu, Y. P. Fang, and Y. H. Xu.

“Preparation of novel Sb2O3/WO3 photocatalysts and their activities under

visible light irradiation”.

In: Materials Research Bulletin 48.6 (2013), pp. 2244–2249.doi: 10.1016/j.materresbull.2013.02.055.

[81] H. Widiyandari, A. Purwanto, R. Balgis, T. Ogi, and K. Okuyama.

“CuO/WO 3 and Pt/WO 3 nanocatalysts for efficient pollutant degradation

using visible light irradiation”.

In: Chemical Engineering Journal 180 (2012), pp. 323–329.doi: 10.1016/j.cej.2011.10.095.

[82] X. Wang, S. Li, H. Yu, J. Yu, and S. Liu. “Ag2O as a New Visible-Light

Photocatalyst: Self-Stability and High Photocatalytic Activity”.

In: Chemistry - A European Journal 17.28 (2011), pp. 7777–7780.doi: 10.1002/chem.201101032.

[83] J.-S. Hu, L.-L. Ren, Y.-G. Guo, et al. “Mass Production and High Photocatalytic

Activity of ZnS Nanoporous Nanoparticles”.

In: Angewandte Chemie 117.8 (2005), pp. 1295–1299.doi: 10.1002/ange.200462057.

[84] A. Koca and M. Sahin. “Photocatalytic hydrogen production by direct sun light

from sulfide/sulfite solution”.

In: International Journal of Hydrogen Energy 27.4 (2002), pp. 363–367.doi: 10.1016/S0360-3199(01)00133-1.

[85] G.-j. Lee and J. J. Wu. “Recent developments in ZnS photocatalysts from

synthesis to photocatalytic applications — A review”.

In: Powder Technology 318 (2017), pp. 8–22.doi: 10.1016/j.powtec.2017.05.022.

[86] M. D. Hernández-Alonso, F. Fresno, S. Suárez, and J. M. Coronado.

“Development of alternative photocatalysts to TiO2: Challenges and

opportunities”. In: Energy & Environmental Science 2.12 (2009), pp. 1231–1257.doi: 10.1039/b907933e.

151

Page 152: Modeling and simulation of photocatalytic degradation of ...

[87] A. Di Paola, E. García-López, G. Marcì, and L. Palmisano.

“A survey of photocatalytic materials for environmental remediation”.

In: Journal of Hazardous Materials 211-212 (2012), pp. 3–29.doi: 10.1016/j.jhazmat.2011.11.050.

[88] M. a. Henderson. “A surface science perspective on TiO2 photocatalysis”.

In: Surface Science Reports 66.6-7 (2011), pp. 185–297.doi: 10.1016/j.surfrep.2011.01.001.

[89] A. Jain, S. P. Ong, G. Hautier, et al. “Commentary: The Materials Project: A

materials genome approach to accelerating materials innovation”.

In: APL Materials 1.1, 011002 (2013). doi: 10.1063/1.4812323.[90] H. Zhang and J. F. Banfield.

“Understanding Polymorphic Phase Transformation Behavior during Growth

of Nanocrystalline Aggregates: Insights from TiO 2”.

In: The Journal of Physical Chemistry B 104.15 (2000), pp. 3481–3487.doi: 10.1021/jp000499j.

[91] X. Chen and S. S. Mao. “Titanium Dioxide Nanomaterials: Synthesis, Properties,

Modifications, and Applications”.

In: Chemical Reviews 107.7 (2007), pp. 2891–2959. doi: 10.1021/cr0500535.[92] G. Li and K. A. Gray. “The solid—solid interface: Explaining the high and unique

photocatalytic reactivity of TiO2-based nanocomposite materials”.

In: Chemical Physics 339.1-3 (2007), pp. 173–187.doi: 10.1016/j.chemphys.2007.05.023.

[93] H. N. Le, F. Babick, K. Kühn, M. T. Nguyen, M. Stintz, and G. Cuniberti. “Impact

of ultrasonic dispersion on the photocatalytic activity of titania aggregates”.

In: Beilstein Journal of Nanotechnology 6.1 (2015), pp. 2423–2430.doi: 10.3762/bjnano.6.250.

[94] European Commission.

“Types and uses of nanomaterials, including safety aspects Accompanying”.

In: CELEX, 52012SC0288 (2012).[95] G. M. Bedinger, J. J. Kohler, and G. J. Wallace. “Titanium - 2015”.

In: U.S. Geological Survey Minerals Yearbook 2015: Volume I, Metals and Minerals.2018. doi: 10.3133/mybvI.

[96] Council of European Union. “Commission regulation (EU) No 1129/2011”.

In: Official Journal of the European Union 54, OJ:L295 (2011), pp. 1–177.doi: 10.3000/19770677.L_2011.295.eng.

152

Page 153: Modeling and simulation of photocatalytic degradation of ...

[97] Scientific Committee on Consumer Safety.

Opinion on Titanium Dioxide (nano form). SCCS/1516/13. 2014.[98] D. S. Hacker and J. B. Butt. “Photocatalysis in a slurry reactor”.

In: Chemical Engineering Science 30.9 (1975), pp. 1149–1158.doi: 10.1016/0009-2509(75)87018-7.

[99] K. M. Lee, C. W. Lai, K. S. Ngai, and J. C. Juan. “Recent developments of zinc

oxide based photocatalyst in water treatment technology: A review”.

In: Water Research 88 (2016), pp. 428–448. doi: 10.1016/j.watres.2015.09.045.[100] D. Moore and Z. L. Wang.

“Growth of anisotropic one-dimensional ZnS nanostructures”.

In: Journal of Materials Chemistry 16.40 (2006), pp. 3898–3905.doi: 10.1039/b607902b.

[101] Research&Markets. “Executive Summary”. In: Zinc Oxide Market by Application& by Region - Global Trends and Forecasts to 2020.Dublin: Research and Markets, 2015.

[102] Umweltbundesamt. “Bekanntmachung der bereits durch die oder auf Grund

der Verwaltungsvorschrift wassergefährdende Stoffe eingestuften Stoffe,

Stoffgruppen und Gemische gemäß § 66 Satz 1 der Verordnung über Anlagen

zum Umgang mit wassergefährdenden Stoffen”.

In: Bundesanzeiger, AT/10.08.2017/B5 (2017).[103] V. Stone, R. Aitken, K. Aschberger, et al. Engineered nanoparticles: Review of

health and environmental safety (ENRHES) - Final Report.Edinburgh Napier University, 2010.

[104] H. Genscher.

“Electrochemical Behavior of Semiconductors under Illumination”.

In: Journal of The Electrochemical Society 113.11 (1966), pp. 1174–1182.doi: 10.1149/1.2423779.

[105] Z. R. Tian, J. A. Voigt, J. Liu, et al. “Complex and oriented ZnO nanostructures”.

In: Nature Materials 2.12 (2003), pp. 821–826. doi: 10.1038/nmat1014.[106] D. M. Schultz and T. P. Yoon.

“Solar Synthesis: Prospects in Visible Light Photocatalysis”.

In: Science 343.6174, 1239176 (2014). doi: 10.1126/science.1239176.[107] ASTM. Standard Tables for Reference Solar Spectral Irradiances: Direct Normal

and Hemispherical on 37° Tilted Surface. G173-03.West Conshohocken, USA: ASTM - International, 2013.

153

Page 154: Modeling and simulation of photocatalytic degradation of ...

[108] S. G. Kumar and L. G. Devi. “Review on Modified TiO2 Photocatalysis under

UV/Visible Light: Selected Results and Related Mechanisms on Interfacial

Charge Carrier Transfer Dynamics”.

In: The Journal of Physical Chemistry A 115.46 (2011), pp. 13211–13241.doi: 10.1021/jp204364a.

[109] M. Pelaez, N. T. Nolan, S. C. Pillai, et al. “A review on the visible light active

titanium dioxide photocatalysts for environmental applications”.

In: Applied Catalysis B: Environmental 125 (2012), pp. 331–349.doi: 10.1016/j.apcatb.2012.05.036.

[110] R. Fagan, D. E. McCormack, D. D. Dionysiou, and S. C. Pillai.

“A review of solar and visible light active TiO2 photocatalysis for treating

bacteria, cyanotoxins and contaminants of emerging concern”.

In: Materials Science in Semiconductor Processing 42 (2016), pp. 2–14.doi: 10.1016/j.mssp.2015.07.052.

[111] L. Wu, J. C. Yu, and X. Fu. “Characterization and photocatalytic mechanism of

nanosized CdS coupled TiO2 nanocrystals under visible light irradiation”.

In: Journal of Molecular Catalysis A: Chemical 244.1-2 (2006), pp. 25–32.doi: 10.1016/j.molcata.2005.08.047.

[112] S. Sato. “Photocatalytic activity of NOx-doped TiO2 in the visible light region”.

In: Chemical Physics Letters 123.1-2 (1986), pp. 126–128.doi: 10.1016/0009-2614(86)87026-9.

[113] R. Asahi, T. Morikawa, T. Ohwaki, and Y. Taga.

“Visible-Light Photocatalysis in Nitrogen-Doped Titanium Oxides”.

In: Science 293.5528 (2001), pp. 269–271. doi: 10.1126/science.1061051.[114] V. Etacheri, M. K. Seery, S. J. Hinder, and S. C. Pillai.

“Highly Visible Light Active TiO(2-x)Nx Heterojunction Photocatalysts”.

In: Chemistry of Materials 22.13 (2010), pp. 3843–3853.doi: 10.1021/cm903260f.

[115] K. Nagaveni, M. S. Hegde, N. Ravishankar, G. N. Subbanna, and G. Madras.

“Synthesis and Structure of Nanocrystalline TiO2 with Lower Band Gap

Showing High Photocatalytic Activity”. In: Langmuir 20.7 (2004), pp. 2900–2907.doi: 10.1021/la035777v.

[116] M. Pelaez, P. Falaras, V. Likodimos, et al.

“Synthesis, structural characterization and evaluation of sol-gel-based NF-TiO2

films with visible light-photoactivation for the removal of microcystin-LR”.

154

Page 155: Modeling and simulation of photocatalytic degradation of ...

In: Applied Catalysis B: Environmental 99.3-4 (2010), pp. 378–387.doi: 10.1016/j.apcatb.2010.06.017.

[117] M. K. Seery, R. George, P. Floris, and S. C. Pillai. “Silver doped titanium dioxide

nanomaterials for enhanced visible light photocatalysis”. In: Journal ofPhotochemistry and Photobiology A: Chemistry 189.2-3 (2007), pp. 258–263.doi: 10.1016/j.jphotochem.2007.02.010.

[118] S. Malato Rodríguez, J. Blanco Gálvez, M. Maldonado Rubio, et al.

“Engineering of solar photocatalytic collectors”.

In: Solar Energy 77.5 (2004), pp. 513–524. doi: 10.1016/j.solener.2004.03.020.[119] S. A. Kalogirou. “Solar thermal collectors and applications”.

In: Progress in Energy and Combustion Science 30.3 (2004), pp. 231–295.doi: 10.1016/j.pecs.2004.02.001.

[120] A. Fernández-García, E. Zarza, L. Valenzuela, and M. Pérez.

“Parabolic-trough solar collectors and their applications”.

In: Renewable and Sustainable Energy Reviews 14.7 (2010), pp. 1695–1721.doi: 10.1016/j.rser.2010.03.012.

[121] M. Tian, Y. Su, H. Zheng, G. Pei, G. Li, and S. Riffat.

“A review on the recent research progress in the compound parabolic

concentrator (CPC) for solar energy applications”. In: Renewable andSustainable Energy Reviews 82.September 2017 (2018), pp. 1272–1296.doi: 10.1016/j.rser.2017.09.050.

[122] Y. Gao, C. Shi, and Y. Wu.

“Luminescence properties of SrB4O7: Eu, Tb phosphors”.

In: Materials Research Bulletin 31.5 (1996), pp. 439–444.doi: 10.1016/S0025-5408(96)00027-X.

[123] F. J. Rivas, F. J. Beltrán, and A. Encinas.

“Removal of emergent contaminants: Integration of ozone and photocatalysis”.

In: Journal of Environmental Management 100 (2012), pp. 10–15.doi: 10.1016/j.jenvman.2012.01.025.

[124] H. J. Round. “A note on carborundum”. In: Electrical World XIX.6 (1907), p. 309.[125] R. Braunstein. “Radiative Transitions in Semiconductors”.

In: Physical Review 99.6 (1955), pp. 1892–1893. doi: 10.1103/PhysRev.99.1892.

[126] N. Holonyak and S. F. Bevacqua.

“Coherent (visible) light emission from Ga(As(1–x)Px) junctions”.

In: Applied Physics Letters 1.4 (1962), pp. 82–83. doi: 10.1063/1.1753706.

155

Page 156: Modeling and simulation of photocatalytic degradation of ...

[127] S. Nakamura, T. Mukai, and M. Senoh. “Candela-class high-brightness

InGaN/AlGaN double-heterostructure blue-light-emitting diodes”.

In: Applied Physics Letters 64.13 (1994), pp. 1687–1689. doi: 10.1063/1.111832.[128] Y. Muramoto, M. Kimura, and S. Nouda. “Development and future of

ultraviolet light-emitting diodes: UV-LED will replace the UV lamp”.

In: Semiconductor Science and Technology 29.8, 084004 (2014).doi: 10.1088/0268-1242/29/8/084004.

[129] UNEP. Minamata Convention on Mercury - text and annexes.Nairobi, Kenya: The United Nations Environment Programme, 2017.

[130] G. Matafonova and V. Batoev.

“Recent advances in application of UV light-emitting diodes for degrading

organic pollutants in water through advanced oxidation processes: A review”.

In: Water Research 132 (2018), pp. 177–189. doi: 10.1016/j.watres.2017.12.079.[131] T.-C. Chao, Y.-T. Lin, C.-Y. Yang, et al. “Highly Efficient UV Organic Light-Emitting

Devices Based on Bi(9,9-diarylfluorene)s”.

In: Advanced Materials 17.8 (2005), pp. 992–996.doi: 10.1002/adma.200401476.

[132] E. Manna, F. Fungura, R. Biswas, J. Shinar, and R. Shinar. “Tunable near UV

microcavity OLED arrays: Characterization and analytical applications”.

In: Advanced Functional Materials 25.8 (2015), pp. 1226–1232.doi: 10.1002/adfm.201403313.

[133] F. Dumur. “Deep-blue Organic Light-emitting Diodes: From Fluorophores to

Phosphors for High-efficiency Devices”.

In: Advanced Surface Engineering Materials.Hoboken, NJ, USA: John Wiley & Sons, Inc., 2016, pp. 561–634.

doi: 10.1002/9781119314196.ch13.

[134] X. Y. Liu, H. T. Ge, Y. Zhao, D. Zhao, J. Fan, and L. S. Liao.

“Novel carbazole derivatives designed by an: Ortho-linkage strategy for

efficient phosphorescent organic light-emitting diodes”.

In: Journal of Materials Chemistry C 6.15 (2018), pp. 4300–4307.doi: 10.1039/c8tc00630j.

[135] ISO Central Secretary.

Determination of the specific surface area of solids by gas adsorption - BET method.ISO 9277:2010(E).

Geneva, Switzerland: International Organization for Standardization, 2010.

156

Page 157: Modeling and simulation of photocatalytic degradation of ...

[136] S. Brunauer, P. H. Emmett, and E. Teller.

“Adsorption of Gases in Multimolecular Layers”.

In: Journal of the American Chemical Society 60.2 (1938), pp. 309–319.doi: 10.1021/ja01269a023.

[137] M. von Ardenne. “Das Elektronen-Rastermikroskop”.

In: Zeitschrift für Physik 109.9-10 (1938), pp. 553–572.doi: 10.1007/BF01341584.

[138] J. Torrent and V. Barrón. “Diffuse Reflectance Spectroscopy”.

In: Methods of Soil Analysis Part 5 – Mineralogical Methods.Madison: SSSA Book, 2008, pp. 367–385. doi: 10.2136/sssabookser5.5.c13.

[139] P. Kubelka and F. Munk. “Ein Beitrag zur Optik der Farbanstriche”.

In: Zeitschrift für technische Physik 11a (1931), pp. 593–620.[140] P. Kubelka.

“New Contributions to the Optics of Intensely Light-Scattering Materials Part I”.

In: Journal of the Optical Society of America 38.5 (1948), p. 448.doi: 10.1364/JOSA.38.000448.

[141] G. Kortüm. Reflectance Spectroscopy. Berlin, Heidelberg: Springer, 1969.doi: 10.1007/978-3-642-88071-1.

[142] K. S. Overway. Environmental Chemistry: An Analytical Approach.Hoboken, New Jersey: John Wiley & Sons, Incorporated, 2017.

[143] D. A. Skoog, S. R. Crouch, and F. Holler. Instrumental Analysis Principles. 7th ed.Boston, USA: Cengage Learning, 2017.

[144] R. Gurke, J. Rossmann, S. Schubert, et al. “Development of a

SPE–HPLC—MS/MS method for the determination of most prescribed

pharmaceuticals and related metabolites in urban sewage samples”.

In: Journal of Chromatography B 990 (2015), pp. 23–30.doi: 10.1016/j.jchromb.2015.03.008.

[145] S. Teixeira, R. Gurke, H. Eckert, K. Kühn, J. Fauler, and G. Cuniberti.

“Photocatalytic degradation of pharmaceuticals present in conventional

treated wastewater by nanoparticle suspensions”.

In: Journal of Environmental Chemical Engineering 4.1 (2016), pp. 287–292.doi: 10.1016/j.jece.2015.10.045.

[146] Ratiopharm. Ciprofloxacin-ratiopharm Filmtabletten (Fachinformation/SPC).2020.

157

Page 158: Modeling and simulation of photocatalytic degradation of ...

[147] J. Rossmann, S. Schubert, R. Gurke, R. Oertel, and W. Kirch.

“Simultaneous determination of most prescribed antibiotics in multiple urban

wastewater by SPE–LC–MS/MS”.

In: Journal of Chromatography B 969 (2014), pp. 162–170.doi: 10.1016/j.jchromb.2014.08.008.

[148] R. H. Howland.

“Methylene Blue: The Long and Winding Road from Stain to Brain: Part 1”.

In: Journal of Psychosocial Nursing and Mental Health Services 54.9 (2016),pp. 21–24. doi: 10.3928/02793695-20160818-01.

[149] R. H. Howland.

“Methylene Blue: The Long and Winding Road From Stain to Brain: Part 2”.

In: Journal of Psychosocial Nursing and Mental Health Services 54.10 (2016),pp. 21–26. doi: 10.3928/02793695-20160920-04.

[150] C. Marx, N. Günther, S. Schubert, et al.

“Mass flow of antibiotics in a wastewater treatment plant focusing on removal

variations due to operational parameters”.

In: Science of The Total Environment 538 (2015), pp. 779–788.doi: 10.1016/j.scitotenv.2015.08.112.

[151] C. R. Wilke and P. Chang.

“Correlation of diffusion coefficients in dilute solutions”.

In: AIChE Journal 1.2 (1955), pp. 264–270. doi: 10.1002/aic.690010222.[152] J. Crank. The mathematics of diffusion. Oxford: Clarendon Press, 1975.[153] J. Liang, H. Edelsbrunner, P. Fu, P. V. Sudhakar, and S. Subramaniam.

“Analytical shape computation of macromolecules: I. molecular area and

volume through alpha shape”.

In: Proteins: Structure, Function, and Genetics 33.1 (1998), pp. 1–17.doi: 10.1002/(SICI)1097-0134(19981001)33:1<1::AID-PROT1>3.0.CO;2-O.

[154] T. A. Halgren. “Merck molecular force field. I. Basis, form, scope,

parameterization, and performance of MMFF94”.

In: Journal of Computational Chemistry 17.5-6 (1996), pp. 490–519.doi: 10.1002/(SICI)1096-987X(199604)17:5/6<490::AID-JCC1>3.0.CO;2-P.

[155] A. Bondi. “Van der waals volumes and radii”.

In: Journal of Physical Chemistry 68.3 (1964), pp. 441–451.doi: 10.1021/j100785a001.

158

Page 159: Modeling and simulation of photocatalytic degradation of ...

[156] R. S. Rowland and R. Taylor.

“Intermolecular Nonbonded Contact Distances in Organic Crystal Structures:

Comparison with Distances Expected from van der Waals Radii”.

In: Journal of Physical Chemistry 100.18 (1996), pp. 7384–7391.doi: 10.1021/jp953141+.

[157] S. L. Tait, Z. Dohnálek, C. T. Campbell, and B. D. Kay.

“n-alkanes on MgO(100). II. Chain length dependence of kinetic desorption

parameters for small n-alkanes.”

In: The Journal of chemical physics 122.16, 164708 (2005).doi: 10.1063/1.1883630.

[158] K. A. Fichthorn, K. E. Becker, and R. A. Miron.

“Molecular simulation of temperature-programmed desorption”.

In: Catalysis Today 123.1-4 (2007), pp. 71–76.doi: 10.1016/j.cattod.2006.12.003.

[159] A. Houas. “Photocatalytic degradation pathway of methylene blue in water”.

In: Applied Catalysis B: Environmental 31.2 (2001), pp. 145–157.doi: 10.1016/S0926-3373(00)00276-9.

[160] F. J. Beltrán, A. Aguinaco, and J. F. García-Araya. “Mechanism and kinetics of

sulfamethoxazole photocatalytic ozonation in water.”

In: Water research 43.5 (2009), pp. 1359–1369.doi: 10.1016/j.watres.2008.12.015.

[161] M. M. Mohamed and M. M. Al-Esaimi.

“Characterization, adsorption and photocatalytic activity of vanadium-doped

TiO2 and sulfated TiO2 (rutile) catalysts: Degradation of methylene blue dye”.

In: Journal of Molecular Catalysis A: Chemical 255.1-2 (2006), pp. 53–61.doi: 10.1016/j.molcata.2006.03.071.

[162] T. Zhang, T. K. Oyama, S. Horikoshi, H. Hidaka, J. Zhao, and N. Serpone.

“Photocatalyzed N-demethylation and degradation of methylene blue in titania

dispersions exposed to concentrated sunlight”.

In: Solar Energy Materials and Solar Cells 73.3 (2002), pp. 287–303.doi: 10.1016/S0927-0248(01)00215-X.

[163] C. Minero, G. Mariella, V. Maurino, and E. Pelizzetti.

“Photocatalytic Transformation of Organic Compounds in the Presence of

Inorganic Anions. 1. Hydroxyl-Mediated and Direct Electron-Transfer

159

Page 160: Modeling and simulation of photocatalytic degradation of ...

Reactions of Phenol on a Titanium Dioxide-Fluoride System”.

In: Langmuir 16.6 (2000), pp. 2632–2641. doi: 10.1021/la9903301.[164] Microprocessor Standards Committee.

IEEE Standard for Floating-Point Arithmetic. IEEE 754-2019.New York, USA: IEEE Standards Association, 2019.

doi: 10.1109/IEEESTD.2019.8766229.

[165] Python Core Team. Python: A dynamic, open source programming language.Python Software Foundation. 2020.

[166] P. Virtanen, R. Gommers, T. E. Oliphant, et al.

“SciPy 1.0: fundamental algorithms for scientific computing in Python”.

In: Nature Methods 17.3 (2020), pp. 261–272. doi: 10.1038/s41592-019-0686-2.[167] S. van der Walt, S. C. Colbert, and G. Varoquaux.

“The NumPy Array: A Structure for Efficient Numerical Computation”.

In: Computing in Science & Engineering 13.2 (2011), pp. 22–30.doi: 10.1109/MCSE.2011.37.

[168] J. D. Hunter. “Matplotlib: A 2D Graphics Environment”.

In: Computing in Science & Engineering 9.3 (2007), pp. 90–95.doi: 10.1109/MCSE.2007.55.

[169] S. van der Walt, J. L. Schönberger, J. Nunez-Iglesias, et al.

“scikit-image: image processing in Python”. In: PeerJ 2.1, e453 (2014).doi: 10.7717/peerj.453.

[170] A. C. Hindmarsh. “ODEPACK, A Systematized Collection of ODE Solvers”.

In: Scientific Computing. Ed. by R. S. Stepleman. Amsterdam, 1983, pp. 55–64.[171] L. Petzold. “Automatic Selection of Methods for Solving Stiff and Nonstiff

Systems of Ordinary Differential Equations”.

In: SIAM Journal on Scientific and Statistical Computing 4.1 (1983), pp. 136–148.doi: 10.1137/0904010.

[172] R. P. Brent. Algorithms for Minimization Without Derivatives.Mineola, New York: Dover Publications. Inc, 2013.

[173] W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery.

Numerical Recipes in Fortran 77 - The Art of Scientific Computing. 2nd ed. 1992,pp. 395–398.

160

Page 161: Modeling and simulation of photocatalytic degradation of ...

[174] K. Levenberg.

“A method for the solution of certain non-linear problems in least squares”.

In: Quarterly of Applied Mathematics 2.2 (1944), pp. 164–168.doi: 10.1090/qam/10666.

[175] D. W. Marquardt.

“An Algorithm for Least-Squares Estimation of Nonlinear Parameters”.

In: Journal of the Society for Industrial and Applied Mathematics 11.2 (1963),pp. 431–441. doi: 10.1137/0111030.

[176] J. J. Moré. “The Levenberg-Marquardt algorithm: Implementation and theory”.

In: Journal of Chemical Information and Modeling. Vol. 53. 9. 1978, pp. 105–116.doi: 10.1007/BFb0067700.

[177] J. Pujol. “The solution of nonlinear inverse problems and the

Levenberg-Marquardt method”. In: GEOPHYSICS 72.4 (2007), W1–W16.doi: 10.1190/1.2732552.

[178] K. J. A. Raj and B. Viswanathan. “Effect of surface area , pore volume and

particle size of P25 titania on the phase transformation of anatase to rutile”.

In: Indian Journal of Chemistry 48A (2009), pp. 1378–1382.[179] H. Chang, C. Su, C. Lo, L. Chen, T. Tsung, and C. Jwo. “Photodecomposition and

Surface Adsorption of Methylene Blue on TiO2 Nanofluid Prepared by ASNSS”.

In: Materials Transactions 45.12 (2004), pp. 3334–3337.doi: 10.2320/matertrans.45.3334.

[180] M. Salaices, B. Serrano, and H. de Lasa.

“Photocatalytic conversion of phenolic compounds in slurry reactors”.

In: Chemical Engineering Science 59.1 (2004), pp. 3–15.doi: 10.1016/j.ces.2003.07.015.

[181] J. Moreira, B. Serrano, A. Ortiz, and H. de Lasa. “A unified kinetic model for

phenol photocatalytic degradation over TiO2 photocatalysts”.

In: Chemical Engineering Science 78 (2012), pp. 186–203.doi: 10.1016/j.ces.2012.04.033.

[182] O. Debono, F. Thévenet, P. Gravejat, et al.

“Gas phase photocatalytic oxidation of decane at ppb levels: Removal kinetics,

reaction intermediates and carbon mass balance”. In: Journal of Photochemistryand Photobiology A: Chemistry 258 (2013), pp. 17–29.doi: 10.1016/j.jphotochem.2013.02.022.

161

Page 162: Modeling and simulation of photocatalytic degradation of ...

[183] A. Hu, X. Zhang, K. D. Oakes, P. Peng, Y. N. Zhou, and M. R. Servos.

“Hydrothermal growth of free standing TiO2 nanowire membranes for

photocatalytic degradation of pharmaceuticals”.

In: Journal of Hazardous Materials 189.1-2 (2011), pp. 278–285.doi: 10.1016/j.jhazmat.2011.02.033.

162


Recommended